Skip to main content

Digital in-line holographic microscopy for label-free identification and tracking of biological cells

Abstract

Digital in-line holographic microscopy (DIHM) is a non-invasive, real-time, label-free technique that captures three-dimensional (3D) positional, orientational, and morphological information from digital holographic images of living biological cells. Unlike conventional microscopies, the DIHM technique enables precise measurements of dynamic behaviors exhibited by living cells within a 3D volume. This review outlines the fundamental principles and comprehensive digital image processing procedures employed in DIHM-based cell tracking methods. In addition, recent applications of DIHM technique for label-free identification and digital tracking of various motile biological cells, including human blood cells, spermatozoa, diseased cells, and unicellular microorganisms, are thoroughly examined. Leveraging artificial intelligence has significantly enhanced both the speed and accuracy of digital image processing for cell tracking and identification. The quantitative data on cell morphology and dynamics captured by DIHM can effectively elucidate the underlying mechanisms governing various microbial behaviors and contribute to the accumulation of diagnostic databases and the development of clinical treatments.

Background

Recent advancements in microscopy technology have facilitated quantitative analysis of cell dynamics in the fields of cell biology and biomedical research [1,2,3,4,5]. The utilization of miniaturized total analytic systems and lab-on-a-chip technologies has increasingly enabled digital imaging techniques to analyze the dynamic behaviors of living cells under various microfluidic conditions [6,7,8,9,10]. For example, optical microscopy has been employed to analyze membrane deformation and rotational motion of erythrocytes flowing in microchannels [11,12,13,14,15,16]. Confocal microscopy was utilized to reconstruct the three-dimensional (3D) morphological structures of various biological samples from their high-resolution stacks of images obtained by point-by-point scanning using a focused illuminating beam [17,18,19,20]. This advanced microscopy technique allowed for the analysis of cell dynamics under different microenvironments. Defocusing microscopy was used to estimate the 3D surface characteristics of cells under various experimental conditions [21,22,23]. However, these microscopy techniques have technical limitations in observing cell dynamics within a wide 3D volume due to their limited depth of field (DOF).

Digital in-line holographic microscopy (DIHM) is a 3D imaging technique that efficiently captures both the 3D positional and morphological information of test samples over time [24,25,26,27,28,29,30,31,32]. Unlike conventional microscopy techniques, which only provide two-dimensional (2D) images on the focal plane of an objective lens, DIHM records consecutive holographic interference signals of test samples that contain 3D volumetric information using a digital image recording device, such as a charge-coupled device camera or a complementary metal-oxide-semiconductor camera. By employing various numerical backpropagation methods, holographic images at different distances from the sensor plane of the DIHM system are reconstructed. Autofocusing algorithms are subsequently applied to determine the 3D positions of test samples and obtain the corresponding in-focus reconstructed images. Four-dimensional (4D; 3D spaces + 1D time) spatio-temporal trajectories of test samples can be extracted from the reconstructed consecutive holographic images. Therefore, the DIHM technique has been widely utilized for precise and quantitative measurement of the 3D behavioral characteristics exhibited by various microscale particles, including flow tracers in microfluidics [33,34,35,36,37,38,39], colloids [40,41,42], microbubbles [43, 44], particulate matter [45, 46], and microorganisms [47,48,49,50].

This review article presents an overview of the fundamental principles and applications of DIHM for quantitative analyses of cell dynamics in 3D volumes. The DIHM-based cell tracking procedures, including DIHM configuration, digital image preprocessing, numerical reconstruction, autofocusing, and particle tracking velocimetry (PTV) algorithms, are summarized. Recent studies on 3D dynamic analysis of living cells using DIHM technique are also discussed, covering a range of organisms such as erythrocytes, spermatozoa, bacteria, dinoflagellates, and algae. Experimental investigations into various dynamic behaviors of living cells, encompassing single-cell motilities, cell-cell interactions, and cell-surface interactions, are reviewed. Finally, recent studies on label-free sensing and classification of different types of living cells, such as diseased cells and microorganisms, demonstrate the potential clinical applications of DIHM technique for facile and accurate diagnosis of cellular diseases.

Principles of DIHM

Optical configuration of DIHM

The basic configuration of DIHM, which uses a point light source, is derived from Gabor holography (Fig. 1a) [24]. It consists of a coherent laser source, a spatial filter, and a digital camera. A coherent laser beam with spherical waves is scattered from a test sample to generate an object beam. When the distance between the light source and the image sensor of the digital camera is sufficiently large, the incident wave can be approximated as a plane wave. A reference beam represents an unaffected wave emitted from the light source. In the DIHM configuration, the object and reference beams propagate in the same direction and interfere to form holographic interference patterns recorded on the image sensor. The magnification ratio can be adjusted by changing the ratio of the distance between the pinhole and the test sample to that between the pinhole and the camera. On the other hand, increasing the distance between the pinhole and the camera decreases the numerical aperture (NA) of the DIHM configuration using spherical waves. In addition, the magnification ratio and light intensity of holographic signals vary depending on the depth-wise position of the test sample. To visualize 3D dynamic behaviors of test samples over a wide volume with extended DOF, a collimated laser beam is utilized to configure the DIHM system (Fig. 1b) [29]. For this purpose, a convex lens converts spherical waves into plane waves, while an additional objective lens attached in front of the image sensor of the camera increases the magnification ratio. The holographic images captured by the DIHM system using plane waves maintain a constant pixel length and resolution, regardless of the depth position of the test sample. This advantage facilitates the design of experimental setups and enables the application of reconstruction algorithms to analyze 3D dynamic behaviors of test samples moving in a large volume. The field of view (FOV) is calculated by dividing the physical dimension of an image sensor by the magnification ratio of the DIHM system. The lateral and longitudinal resolutions are defined as |Δrlateral|≥ 0.5λ/NA and |Δrlongitudinal|≥ 0.5λ/NA2, where λ represents the wavelength of the light source [26].

Fig. 1
figure 1

Schematics of the optical configurations of digital in-line holographic microscopy using a point source (a) and a collimated beam (b)

Several other in-line holographic configurations have recently been introduced. The DIHM configuration can be further simplified by replacing the coherent laser source with an inexpensive partially coherent light source, such as a light-emitting diode [32, 51,52,53]. Due to the lack of expensive objective lenses and a coherent laser source, the lensless DIHM system can be employed for developing compact portable devices to monitor various microscale particulates. Instead of using a coherent light source, spatially incoherent sunlight or fluorescence has recently been utilized for developing incoherent digital holography techniques, including optical scanning holography, Fresnel incoherent correlation holography, and coded aperture correlation holography [54,55,56,57]. To resolve the twin-image problem of DIHM configuration, single-shot in-line phase-shifting interferometry has been developed [58,59,60]. A phase-shifting array device is employed to capture multiple holograms with various phase shifts. This enables precise measurements of the phase information of a test object from the captured holograms and enhances the quality of reconstructed images. Multi-wavelength DIHM systems are utilized to extract multiple single-wavelength holograms from one single-shot hologram [61,62,63,64]. These extracted holograms are then used for phase unwrapping and comprehensive analysis of 3D morphological characteristics of biological samples illuminated with different wavelengths.

Preprocessing of holographic images

Raw holographic images of a test sample contain unintended background noises induced by static dust particles and scratches on optical components. There are several methods available to remove these background noises and acquire clean holographic signals from the test sample. Firstly, the raw holographic images are normalized using the background image obtained by recording a holographic image without the test sample [65,66,67]. Secondly, normalization is performed based on the illumination intensity to effectively suppress multiplicative artifacts [40, 68]. Thirdly, spatially invariant background noises are subtracted from the raw holographic images by calculating an ensemble average of hundreds of consecutive holograms [29, 69]. Further improvement of the signal-to-noise ratio in holographic images can be achieved by employing various denoising techniques, such as band-pass filters [70,71,72], mean filters [73], median filters [74], Wiener filters [75], local-mean-subtraction filters [76], spectral filters [77], wavelet-based denoising [78], non-local means filtering [79], correlation-based denoising [80], and deep-learning-based denoising methods [81,82,83]. Super-resolution techniques can also be adopted to achieve high-resolution holographic imaging results [84,85,86,87,88].

Numerical reconstruction of holographic images

Holographic images located at different depths away from the image plane can be numerically reconstructed by adopting several light diffraction theories [89, 90]. Kirchhoff-Helmholtz transform has been utilized to reconstruct holographic images (Hr) from the original holographic image (Ho) recorded by a DIHM setup with a point light source [91,92,93]. The reconstruction equation of the Kirchhoff-Helmholtz transform is expressed as follows:

$$H_{\mathrm r}(r)=\frac1{4\pi}\int_s{d\psi H_{\mathrm o}(\psi)}\exp\left[ik\psi\cdot\left(\frac{r}{|\psi|}\right)\right]\,$$
(1)

where Ho(ψ) is the contrast image on the detector screen S at coordinates ψ = (x, y, l) located at a depth-wise distance l away from the point source. r and k denote the position vector from the point source and the wave number of the light source, respectively.

The Kirchhoff’s theory yields precise empirical results. However, the boundary conditions of Kirchhoff’s theory imply the absence of waves behind the aperture, leading to mathematical and physical contradictions [89]. To eliminate these inconsistencies associated with the boundary conditions in the Kirchhoff’s theory, the Rayleigh-Sommerfeld diffraction integral is utilized to reconstruct holograms of both spherical and plane waves [94,95,96,97]. The Rayleigh-Sommerfeld propagator (h), which enables wave field recovery, can be expressed as follows:

$$h = \frac{1}{2\pi }\frac{\partial }{\partial z}\frac{\exp (ikr)}{r} = \,\frac{1}{2\pi }\frac{\partial }{\partial z}\frac{{\exp \left[ {ik\sqrt {{{(\xi - x)}^2} + {{(\eta - y)}^2} + {z^2}} } \right]}}{{\sqrt {{{(\xi - x)}^2} + {{(\eta - y)}^2} + {z^2}} }}\,$$
(2)

where x and y, and ξ and η are the spatial coordinates on the image and reconstruction planes, respectively. z is the distance from the image plane to the reconstruction plane. The diffraction integral is expressed as follows:

$$H_{\mathrm r}(\xi,\eta;z)=\iint H_{\mathrm o}(x,y;0)\cdot h(\xi-x,\eta-y;z)dxdy$$
(3)

where Hr(ξ,η;z) and Ho(x,y;0) are the reconstructed and original holographic images, respectively.

Based on the Fresnel approximation, the diffraction integral can be converted into a simpler expression [27, 80, 98]. The hologram reconstruction process employing the Fresnel transformation can be expressed as follows:

$$H_{\mathrm r}(\xi,\eta;z)=\frac{\exp(ikz)}{i\lambda z}\iint H_{\mathrm o}(x,y;0)\cdot\exp\left\{\frac{ik}{2z}\left[{(\xi-x)}^2+{(\eta-y)}^2\right]\right\}dxdy\,$$
(4)

The Fresnel approximation exhibits precise reconstruction performance for small diffraction angles. The Fraunhofer approximation can be adopted to further simplify the transformation equation in the following manner:

$$H_{\mathrm r}(\xi,\eta;z)=\frac{\exp(ikz)\exp\left[\frac{ik}{2z}\left(\xi^2+\eta^2\right)\right]}{i\lambda z}\iint H_{\mathrm o}(x,y;0)\cdot\exp\left[-\frac{ik}z\left(\xi x+\eta y\right)\right]dxdy$$
(5)

where the quadratic terms of x2 + y2 are omitted [89]. The Fraunhofer transformation facilitates rapid calculation of the propagating wavefronts in far-field imaging.

Angular spectrum method is usually employed for reconstructing holographic images captured by a DIHM setup using plane waves [29, 99]. The angular spectrum of the wavefront recorded in the original holographic image is obtained through the application of fast Fourier transform (FFT), enabling extraction of the spatial frequency components distribution contained in the holographic image. Each spatial frequency component propagates through the space at different distances and angles. By applying the inverse FFT to these propagated spatial frequency components, a new hologram reconstructed at a depth-wise distance z from the original hologram can be obtained. This method does not require the minimum z-distance or any assumptions such as the Fresnel approximation, making it suitable for conducting 3D dynamic analyses of test samples moving in a large volume. The reconstruction equation of the angular spectrum method is expressed as follows:

$$H_{\mathrm r}(\xi,\eta;z)=F^{-1}\left\{F\left[H_{\mathrm o}(x,y;0)\right]\text{exp}\left[ikz\sqrt{1-\left(\lambda f_x\right)^2-\left(\lambda f_y\right)^2}\right]\right\}\,$$
(6)

where F and F−1 represent FFT and inverse FFT, respectively. fx and fy denote the spatial frequencies of x and y coordinates, respectively. To further improve the spatial resolution and signal-to-noise ratio of the reconstructed holograms, additional methodologies are adopted to suppress the twin-image problem associated with DIHM technique [61, 100,101,102,103,104,105,106].

3D localization of test samples using autofocusing algorithms

The 3D positional information of test samples is determined based on the reconstructed holograms, which are obtained by numerically reconstructing holograms at different depth-wise distances from an original hologram of a test sample. These reconstructed holograms are then projected into a single image plane. In-plane (x, y) position of each particle recorded on the projected hologram is determined by identifying extreme values in local intensity, image contrast, or sharpness. Among the reconstructed holograms of the object at the determined in-plane position, the degree of sharpness and image contrast (i.e., focus value) of the holograms is quantified by adopting various autofocusing functions. Subsequently, the depth-wise position z of each particle is determined by searching an extreme peak in focus values obtained by using focus functions, such as gradient (GRA), Laplacian (LAP), weighted spectral (SPEC), Tamura coefficient (TC), and variance (VAR) focus functions. These focus functions used for calculating focus values are defined as follows:

$$GRA(z) = \sum\limits_{\xi ,\eta } {\left| {\nabla {H_\text r}(\xi ,\eta ;z)} \right|}$$
(7)
$$LAP(z) = \sum\limits_{\xi ,\eta } {{{\left[ {{\nabla^2}{H_\text r}(\xi ,\eta ;z)} \right]}^2}}$$
(8)
$$SPEC(z)=\sum\limits_{f_\xi,f_\eta}\log\left\{1+\left|F\left[H_{\mathrm r}(\xi,\eta;z)-{\overline H}_\text{r}(z)\right]\right|\right\}$$
(9)
$$TC(z) = \sqrt {\frac{{\sigma \left[ {{H_\text r}(\xi ,\eta ;z)} \right]}}{{{{\overline H}_\text r}(z)}}}$$
(10)
$$VAR(z) = \frac{1}{{{N_\xi }{N_\eta }}}\sum\limits_{\xi ,\eta } {{{\left[ {{H_\text r}(\xi ,\eta ;z) - {{\overline H}_\text r}(z)} \right]}^2}}$$
(11)

where σ is the standard deviation [107, 108]. The resolution of the reconstructed hologram is Nξ × Nη pixels. fξ and fη denote the spatial frequencies of ξ and η coordinates, respectively. \({\overline H_r}(z)\) represents the spatial average value of a hologram reconstructed at the depth-wise position z. In the Rayleigh-Sommerfeld diffraction integral-based reconstruction process, the Gouy phase anomaly is utilized for the 3D localization of test samples [62, 97, 109]. Additionally, Gini’s index [110], Tenengrad function [111], Brenner function [112], DarkFocus algorithm [113], spectral L1 norm [114, 115], and novel deep-learning-based methodologies [116,117,118,119,120] are employed for autofocusing. Due to variations in experimental conditions such as the size and shape of test samples and the relative refractive index of media, it is important to consider an appropriate autofocusing method for a given experimental condition.

3D PTV algorithms for cell tracking

The 3D dynamic behaviors of test samples can be analyzed by extracting the 3D positional information from reconstructed consecutive holographic images. Furthermore, the trajectory of test samples can be easily obtained by connecting the 3D positions of individual objects in subsequent holograms. Several 3D PTV algorithms can be applied to thousands of 3D positional information of test samples to obtain their trajectories. For example, the two-frame PTV algorithm based on iterative estimation of match probability can be utilized for rapid tracking of particle trajectory using only two successive image frames [121]. The Crocker-Grier algorithm can be employed to search for the probable set of particle indices and locations in each frame among the successive and preceding image frames, taking into account various factors such as size, intensity, and displacement of particles to ensure accurate tracking [122, 123]. Additionally, high-order multi-frame tracking algorithms are employed for robust and accurate 3D Lagrangian tracking of particles [124,125,126,127,128]. Machine learning-based cell tracking algorithms can also be adopted to improve the performance of holographic PTV measurements. A neural network is used for nonlinear global regression to filter out random noises present in PTV data and reconstruct the entire flow field from captured photographs [129]. To overcome technical limitations associated with previous PTV algorithms when dealing with highly concentrated tracer particles and high-speed flows, a long short-term memory network is utilized to predict the subsequent velocity of a tracer particle based on its past PTV data [130].

Alternative holographic processing techniques

Conventional digital image processing methods for holographic PTV typically include preprocessing, numerical reconstruction, autofocusing, and particle tracking. Recently, several alternative holographic processing techniques deviating from traditional categories have been proposed. For example, 3D volumetric deconvolution method utilizes a point-spread function to enhance the optical features contained in reconstructed holograms [131,132,133]. It effectively resolves the superimposed out-of-focus signals of highly concentrated particles. Additionally, various inverse reconstruction methods employing fused lasso regularization [134], Tikhonov regularization [135], and iterative predictive algorithm [136] are introduced to overcome the technical limitations of DIHM in terms of particle concentration and reconstruction dynamic range.

Recently, artificial intelligence (AI) has been applied to expedite the computational time required for numerical reconstruction and autofocusing procedures in digital image processing. Specifically, a convolutional neural network (CNN) is trained using holograms of test samples and corresponding ground-truth depth-wise position labels [118]. By leveraging this trained CNN model, it can directly predict the depth-wise positions of test samples from their holograms without resorting to numerical reconstruction and autofocusing procedures. To achieve this, a CNN architecture composed of convolutional layers, up-sampling blocks, and nonlinear activation functions is trained with pairs of raw holograms and their corresponding reconstructed amplitude and phase maps [104]. Additionally, a U-Net architecture is utilized to predict 3D locations of highly concentrated particles based on input holograms, depth maps, and maximum phase projections [137]. Furthermore, a fusion approach involving two U-Nets is trained with raw holograms at the input layer of one down-sampling path and pairs of intensity and phase maps at the output layer of two up-sampling paths [138]. Moreover, a generative adversarial network, which utilizes mutual training of the generator and discriminator, is employed to generate in-focus intensity and phase maps from an input hologram [139,140,141]. In addition to these techniques mentioned above, a Fourier imager network consisting of spatial Fourier transform modules can provide a global receptive field for processing holographic diffraction patterns obtained from test samples [142]. Lastly, a self-supervised learning model is trained by using a physics-consistency loss along with synthetic images instead of generating experimental datasets [143]. Therefore, recent advancements in AI techniques facilitate rapid hologram reconstruction and precise localization in cell tracking, thereby replacing the time-consuming digital image processing procedures of traditional DIHM methods.

Applications of DIHM to various microscale biological cells

4D tracking of human blood cells

The DIHM technique enables effective analysis of 4D dynamic behaviors of human blood cells, including neutrophils and erythrocytes (Table 1) [144,145,146,147,148,149,150,151]. For example, rapid movement of neutrophils (HL60 cells) was visualized with lateral and longitudinal resolutions in the range of a few micrometers [144]. A DIHM system was employed to track cell migration of unlabeled asthmatic and non-asthmatic neutrophils at a high temporal resolution [145]. Their study comparatively investigated the averages of undirected speed and outward velocity of neutrophils exposed to various chemotactic stimuli (interleukin-8 and N-formylmethionyl-leucyl-phenylalanine) under different mechanical environments such as collagen stiffness and pore size. The 3D velocity profile of erythrocytes in Hagen-Poiseuille flows was measured depending on their radial position while flowing through a microtube with an inner diameter of 350 μm [146]. Statistical analysis was conducted on the inertial migration phenomena exhibited by erythrocytes moving in cylindrical and rectangular microchannels under various microfluidic conditions [147, 148]. 3D spatial distributions of erythrocytes and microspheres were investigated under different shear rates. Quantitative comparisons were made between healthy and hardened erythrocytes, regarding their lateral migration, deformation index, and orientation for various flow rates of viscoelastic fluids in a rectangular microchannel (Fig. 2a) [149]. The typical experimental setup for microfluidic DIHM measurements is illustrated in Fig. 2ai. Digital image processing procedures involving background subtraction (Fig. 2aii), depth localization using a TC focus function (Fig. 2aiii), as well as in-plane positioning (Fig. 2aiv), were utilized to detect the 3D spatial distribution of erythrocytes. The spatial distributions of spherical particles, hardened erythrocytes, and normal erythrocytes in microchannels were obtained with varying flow rates in microchannels (Fig. 2av-vii).

Table 1 Summary of previous studies on four-dimensional (4D) tracking of human blood cells using digital in-line holographic microscopy
Fig. 2
figure 2

Applications of digital in-line holographic microscopy (DIHM) to track human erythrocytes and spermatozoa. a Lateral migration of hardened and normal erythrocytes in viscoelastic flows under different microfluidic conditions. Experimental setup for the microfluidic measurement (i). Digital image processing procedure: background subtraction (ii, scale bar = 10 μm), depth localization using a Tamura coefficient (TC) focus function (iii, scale bar = 10 μm), in-plane positioning (iv, scale bar = 10 μm), and 3D spatial distributions of spherical particles, hardened erythrocytes, and normal erythrocytes measured using DIHM (v-vii). Reprinted from ref. [149], Copyright 2017. b Measurement of 3D locations and orientations of erythrocytes using DIHM and deep learning techniques. Digital image processing procedure: raw hologram (i, scale bar = 20 μm), background subtraction (ii, scale bar = 20 μm), projection (iii, scale bar = 20 μm), depth localization using a gradient focus function (iv, scale bar = 5 μm), in-plane angle measurement (v, scale bar = 5 μm), and 3D positions and orientations of erythrocytes measured using DIHM (vi). Reprinted with permission from ref. [151], Copyright 2023, Elsevier B.V. c Transitions between different swimming patterns of a human spermatozoon. Hyper-activated (i, iv) and helical patterns (iii) are observed in a whole trajectory of the human spermatozoon (ii). Reprinted from ref. [152], Copyright 2012

Stereo-holographic microscopy was developed to obtain 3D positional, orientational, and morphological information on non-spherical particles, including ellipsoidal particles and erythrocytes [150]. To accurately measure the 3D orientations of erythrocytes, a deep-learning-based DIHM technique was developed, which can estimate the in-plane and out-of-plane angles of erythrocytes from their in-focus reconstructed holograms (Fig. 2b) [151]. The signal-to-noise ratio of holographic signals of three erythrocytes recorded in a raw hologram (Fig. 2bi) was enhanced by subtracting background noises (Fig. 2bii). The reconstructed holograms of erythrocytes were projected on a single image plane to determine in-plane locations of erythrocytes by searching for local intensity extrema (Fig. 2biii). Depth locations and in-focus reconstructed holograms of erythrocytes were obtained by employing a GRA focus function (Fig. 2biv). The in-plane angle (φ) of each erythrocyte was determined by evaluating the inclined angle of its major axis (Fig. 2bv). A convolutional autoencoder was utilized to increase the number of holographic images with respect to the out-of-plane angle of erythrocytes. Subsequently, a CNN for regression analysis was used to directly predict the out-of-plane angle (θ) from the reconstructed hologram of each erythrocyte (Fig. 2bvi). On the other hand, previous studies on erythrocytes have mainly focused on measuring their 3D translational dynamics in straight microchannels. Further development is required for DIHM technique with the aid of AI to analyze the 3D rotational dynamics exhibited by moving erythrocytes within more complex microfluidic conduits.

4D tracking of spermatozoa

The swimming motility of human and animal spermatozoa can be analyzed by visualizing 3D trajectories of spermatozoa swimming in a 3D volume using a DIHM technique (Table 2) [152,153,154,155,156,157]. For example, submicron accuracy was achieved in tracking various swimming patterns of human spermatozoa, including typical, helical, hyper-activated, and hyper-helical patterns [152]. The transitions between swimming patterns within each sperm’s trajectory were statistically investigated using a lens-free imaging platform (Fig. 2c). A whole trajectory of a human spermatozoon (Fig. 2cii) includes transitions from the hyper-activated pattern (Fig. 2civ) to the helical pattern (Fig. 2ciii), and back to the hyper-activated pattern again (Fig. 2ci). By adopting the lensless DIHM configuration, a cost-effective microscopic device can be established for analyzing the swimming motions of human spermatozoa with high resolution and a large FOV [153]. Spiral trajectories of goat spermatozoa exhibited significantly higher concentration and intense motility compared to those of human spermatozoa [154]. Quantitative visualization was performed on the 3D trajectories of Arbacia punctulata spermatozoa which navigate through 3D chemoattractant gradients provided by an egg for fertilization, revealing their tracking process toward the egg [155]. The swimming motions of horse spermatozoa were categorized into six different patterns, including irregular, linear, planar, helical, ribbon, and hyper-progressive patterns [156]. Free-swimming spermatozoa of normal and unhealthy mice were also comparatively investigated [157]. A double-knockout mouse model lacking tubulin glycosylation was generated by targeting the initiating glycylases tubulin-tyrosine ligase-like (TTLL) 3 and TTLL8 (Ttll3−/−Ttll8−/− mouse). The DIHM technique was employed to measure the 3D behaviors of spermatozoa from both a normal mouse and a Ttll3−/−Ttll8−/− mouse. The spermatozoa from the normal mouse exhibited twisted ribbon patterns, while those from the Ttll3−/−Ttll8−/− mouse displayed helical patterns. As the spermatozoa from the Ttll3−/−Ttll8−/− mouse approached the wall of the observation chamber, their swimming patterns transitioned from helical to circular motion. Consequently, this kind of change in swimming patterns disrupted their progressive movement. These experimental studies show that the DIHM technique possesses sufficient resolution and measurement accuracy for assessing the 3D swimming motions of spermatozoa.

Table 2 Summary of previous studies on four-dimensional (4D) tracking of spermatozoa using digital in-line holographic microscopy

4D tracking of bacteria

The free-swimming behaviors of various bacteria have been quantitatively analyzed using a DIHM technique to reveal the underlying mechanisms of microbial motility (Table 3) [62, 80, 158,159,160,161,162,163,164,165,166,167,168,169,170,171]. Pseudomonas aeruginosa (P. aeruginosa) exhibited several swimming behaviors, including meandering, oscillation, helix, pseudohelix, and twisting patterns, and the transitions between different patterns were analyzed (Fig. 3a) [158]. Statistical comparisons were made on the 3D swimming speed and turning angle distributions of P. aeruginosa, Agrobacterium tumefaciens, and Escherichia coli (E. coli) [159]. Submicron-scale kinematics, morphological shape, and orientation measurements were conducted on individual E. coli to assess their motilities, including swimming speed, tumbling motion, and wobbling motion [62, 80, 160]. The body-angle rotation during runs, tumbles, and pole reversal in E. coli was measured using a C-implemented discrete dipole approximation code and the Levenberg-Marquardt algorithm [161]. The filament compositions of Shewanella putrefaciens were observed to affect the flagellar morphology and free-swimming trajectories [162].

Table 3 Summary of previous studies on four-dimensional (4D) tracking of bacteria using digital in-line holographic microscopy
Fig. 3
figure 3

Applications of digital in-line holographic microscopy to track various unicellular microorganisms. a Trajectories of swimming Pseudomonas aeruginosa obtained by using DIHM (i, ii) and the corresponding statistical analysis of various swimming patterns, including meander, oscillation, helix, pseudohelix, and twisting patterns (iii, iv). Reprinted from ref. [158], Copyright 2014. b Various swimming patterns of Escherichia coli (E. coli) in near-surface and bulk regions. Trajectories of swimming E. coli (i, viii). Swimming patterns of E. coli in the bulk region: running and tumbling motions (ii) and slow random walk (iii). Swimming patterns of E. coli in the near-surface region: gyrating on a surface (iv), attaching and detaching motions (v), running and tumbling motions (vi), and swimming in circles (vii). Reprinted with permission from ref. [163], Copyright 2014, American Physical Society. c 3D trajectories of solitary and chain-forming Cochlodinium polykrikoides. Reprinted with permission from ref. [172], Copyright 2010, Springer-Verlag. d Trajectories of Prorocentrum minimum in helical motions (i-iii), obtained using DIHM. Probability density functions (PDFs) of helix parameters in the near and bulk regions: radius (R, iv) and pitch (P, v). Statistical differences in helix parameters between the near and bulk regions, represented as probability values (P-values): curvature (κ, vi) and torsion (τ, vii). Reprinted with permission from ref. [173], Copyright 2016, Springer-Verlag Berlin Heidelberg

Quantitative measurement of bacteria cell-surface interactions is essential to understand diverse microbial behaviors, such as bacterial adhesion and biofilm formation, and the development of biomedical and antibiofouling surfaces. Various swimming motions of E. coli were observed in both near-surface and bulk regions (Fig. 3bi, viii), encompassing running and tumbling (Fig. 3bii, vi), slow random walk (Fig. 3biii), gyrating on a surface (Fig. 3biv), attaching and detaching (Fig. 3bv), and circular swimming (Fig. 3bvii) [163]. In the near-surface region, tumbling motions of E. coli were reduced by 50%, while reorientations were restricted to surface-parallel directions, impeding their escape from this region. Shear flow generated in the near-surface area promoted tumbling and reorientation movements, thereby enhancing bacterial dispersion [164]. Surface hydrophobicity reduced the swimming speed of E. coli in the near-surface region, promoting their landing on and adhesion to the surface [165].

The wall entrapment mechanism of E. coli was investigated by quantifying various parameters such as cell axis ratio, vertical speed, collision angle, pitch angle, and wobbling angle [166]. 3D motion analysis of E. coli and Pseudomonas species over biodegradable poly(ε-caprolactone)-based polymers revealed that enzymatic degradation rate was inversely correlated with irreversible adhesion [167]. Similarly, decreasing surface stiffness led to reduced bacterial adhesion for E. coli and Pseudomonas species on polydimethylsiloxane surfaces [168]. Swimming behaviors were found to differ between wild-type strains and isogenic flagellar stator mutants of P. aeruginosa in near-surface environments [169]. Shewanella species exhibited faster swimming speeds and longer trajectories in the bulk region, despite accumulating in the region near the surface [170]. Enterobacter sakazakii (E. sakazakii) showed adaptive swimming behaviors in the region near the surface coated with sessile probiotics, reducing wall accumulation [171].

4D tracking of dinoflagellates

Swimming speeds of Alexandrium ostenfeldii, Alexandrium minutum, and Alexandrium tamarense were compared at different temperatures [174]. Prey-induced changes in the swimming behaviors of Karlodinium veneficum (K. veneficum) and Pfiesteria piscicida were compared in terms of the radius and pitch of their helical swimming trajectories, as well as their translational and angular velocities [98]. Quantitative visualization was performed on the 3D trajectories of predatory K. veneficum and prey Storeatula major immobilized by karlotoxins [175]. Using a DIHM technique, the helical swimming trajectories of Cochlodinium polykrikoides (C. polykrikoides) and Prorocentrum minimum (P. minimum) were also analyzed [72]. The motile characteristics of solitary cells vs. chain-forming cells showed that the swimming speed, helix radius, and pitch of 3D trajectories increased with an increasing number of cells in the C. polykrikoides chain (Fig. 3c) [172]. As the viscosity of the surrounding media increased, both the swimming speed and flagella beating frequency decreased for P. minimum [176]. The hydrodynamic power consumed for the swimming motion of P. minimum was quantitatively estimated. The analyzed 3D swimming trajectories and helix parameters indicated that motility and thrust generation were higher in the near-surface region for P. minimum (Fig. 3d) [173]. Based on the measured 3D trajectories of P. minimum showing helical motions (Fig. 3di-iii), the basic helix parameters, including radius (R, Fig. 3div), pitch (P, Fig. 3dv), curvature (κ, Fig. 3dvi), and torsion (τ, Fig. 3dvii), were evaluated in the near and bulk regions. An abrupt drifting motion exhibited by P. minimum was tracked using a lens-free configuration [177]. Previous studies on the 4D tracking of dinoflagellates are summarized in Table 4 [72, 98, 172,173,174,175,176,177].

Table 4 Summary of previous studies on four-dimensional (4D) tracking of dinoflagellates using digital in-line holographic microscopy

4D tracking of other biological cells

3D trajectories of free-swimming algae Tetraselmis species in seawater were measured using a DIHM technique [178]. Unknown species of algae and bacteria were collected from glacial meltwater and investigated using DIHM in terrestrial and exobiological studies [179]. The relationship between inertial migration and elastic shell compliance was analyzed by measuring the spatial distributions of normal and hardened Chlorella cells in a pipe flow [180]. 3D swimming trajectories of Dunaliella primolecta (D. primolecta) in shear flow were measured to investigate the shear-induced algal migration, compared to those in a quiescent fluid [181]. Complex helical trajectories and velocity fluctuations of free-swimming D. primolecta across different growth phases were also statistically measured using the DIHM technique [134, 182]. Ulva zoospores exhibited various swimming patterns in the near-surface and bulk regions, including straight path, gyration, search circle, orientation, and wobbling motions [183]. Swimming velocity and diving direction of Ulva zoospores near glass surfaces were statistically analyzed [184]. Additionally, 3D trajectories of Ulva zoospores were monitored over various surfaces, such as polyethylene glycol coating, acid-washed glass, and tridecafluorooctyl-triethoxysilane-coated glass surfaces [185].

Underwater DIHM was utilized to observe a diverse range of microorganisms in oceans or lakes, such as Paramecium species, Ciliate species, Didinium species, and Coscinodiscus wailesii (diatom) [26]. The 3D motility of blood-parasite Trypanosoma brucei was quantified under various external conditions to investigate the mechanism of immune evasion [186]. Optimal chemotactic behaviors of Haloarchaea (Haloferax species and Haloarcula species) were examined to reveal the survival strategies of archaea under extreme, nutrient-poor conditions [187]. The 3D trajectory analysis unveiled the active diving motion exhibited by a neuroblastoma cell [177]. Infective and non-infective intracellular parasites (Leishmania mexicana) displayed distinctive swimming patterns and chemotaxis towards human cells [188]. Furthermore, DIHM technique was utilized for measuring size distributions and trajectories of airborne virus-laden droplets and aerosols, such as MS2 bacteriophage and severe acute respiratory syndrome coronavirus 2 [189]. These previous studies underscore the immense potential of DIHM technique for in situ, real-time, and non-invasive measurements of 3D dynamic behaviors across various biological cells. A summary table containing these studies on 4D tracking is presented in Table 5 [26, 134, 177,178,179,180,181,182,183,184,185,186,187,188,189].

Table 5 Summary of previous studies on four-dimensional (4D) tracking of other biological cells using digital in-line holographic microscopy

Label-free identification of biological cells

Holographic images of biological cells can be usefully utilized for label-free cell identification. A lens-free holographic imaging platform was developed for on-chip cytometry to automatically characterize erythrocytes, yeast cells, E. coli, and micro-particles of various sizes [190]. A custom-made decision algorithm was introduced to match the detected hologram textures of arbitrary objects with hologram datasets. The lens-free holographic on-chip imaging platform was employed for enumeration and volume measurement of human blood cells, as well as differentiation between various types of white blood cells, including granulocytes, monocytes, and lymphocytes [191]. The three different types of unlabeled leukocytes were classified by evaluating their cellular size and internal complexity [192]. The focal lengths of real focus and virtual focus were determined from light-intensity profiles obtained from reconstructed holographic images of erythrocytes [193]. The real focal length of erythrocytes decreased after the morphological transition from discocytes to echinocytes and spherocytes with the lapse of duration time of blood storage. Breast cancer cells and ovarian cancer cells were enumerated and characterized based on the in-focus scattered light intensity and cell diameter obtained from their holographic images [194]. Bacterial colonies of E. coli and Staphylococcus intermedius were classified using principal components analysis applied to optical signatures, such as reconstructed amplitude and phase maps [195]. The growth characteristics of Haematococcus pluvialis were experimentally analyzed by measuring the variation in cell size under different levels of light stress [196].

With the rapid advancements of AI, a variety of machine learning and deep learning algorithms have been employed for label-free cell classification. A decision tree was employed to classify 3 distinct types of erythrocytes, including discocytes, echinocytes, and spherocytes [197]. To train a machine learning algorithm, numerous features of erythrocytes were quantified, including morphological traits, intensity distributions of holograms, and optical focusing characteristics. The classification of healthy and unstained malaria-infected erythrocytes was achieved by using a support vector machine (SVM) algorithm trained with morphological and light scattering characteristics [198]. Similarly, the SVM algorithm with a linear kernel trained on various features such as basic morphologies, optical characteristics, and translational and rotational invariants was used to classify different types of leukocytes [199].

Label-free classification techniques were developed using a decision tree algorithm trained with characteristic metrics of cell size and intensity values of holograms to enumerate erythrocytes, peripheral blood mononuclear cells, and breast cancer cells [200]. A CNN consisting of 5 convolutional layers was utilized to classify human mammary gland epithelial cells, breast cancer cells, and esophageal cancer cells [201]. In-flow enumeration of breast cancer cells and ovarian cancer cells from lysed blood samples containing white blood cells was performed using a custom-built shallow network [202]. The viability and concentration of yeast cells were evaluated by employing the SVM algorithm trained with spatial features extracted from the reconstructed amplitude and phase maps [203]. A deep-learning-based architecture named You Only Look Once version 5 was employed to directly predict the viability of yeast cells from denoised holograms without a hologram reconstruction process [204]. Diatom phytoplankton, diatom pennate, Navicula species, and Selenastrum species were classified by a random forest algorithm trained with various features, such as optical volume, coefficient of variation, mean optical path length, projected area, cell skewness, and cell kurtosis [205]. The death rate of algal cells in the East China Sea was assessed using the SVM algorithm trained with features collected from reconstructed amplitude and phase maps obtained from holographic images of Prorocentrum lima algae [206]. A 3D CNN model was utilized to measure the number of clustered algae Phaeodactylum tricornutum [207]. With the assistance of AI-based DIHM technique, most biological cells can be detected and enumerated with high throughput. These studies on AI-based label-free identification methods for biological cell analysis are summarized in Table 6 [197,198,199,200,201,202,203,204,205,206,207].

Table 6 Summary of previous studies on artificial intelligence (AI)-based label-free identification of biological cells using digital in-line holographic microscopy

Comparison with other 3D imaging techniques

Comparison with off-axis digital holographic microscopy (DHM)

Off-axis DHM has been extensively investigated for measuring 3D phase information in various biological samples [28]. In the off-axis DHM systems, there exists a slight difference in the propagating directions of reference and object waves. A band-pass filter is commonly employed to separate real, twin, and zero-order images in the frequency domain of a recorded hologram. The off-axis DHM systems have been utilized for quantitative phase imaging [208,209,210], holographic tomography [211, 212], and dynamic analysis of biological samples [213,214,215,216]. On the contrary, obtaining clear real images from holograms recorded by DIHM systems is challenging due to the twin-image problem. To address this issue during the hologram reconstruction process, iterative phase retrieval methods [100, 101, 105] and deep learning techniques [104, 217, 218] have been developed. However, sparse test samples with weak phase fluctuations are prerequisites to ensure high precision in phase retrieval using DIHM systems. Therefore, off-axis DHM is suitable for analyzing detailed 3D morphology of biological samples.

On the other hand, the space bandwidth product (SBP) of reconstructed holograms obtained from an off-axis DHM system is somewhat limited compared to DIHM [219]. The SBP of reconstructed holograms is determined by multiplying the FOV and spatial frequency bandwidth. In comparison, the SBP values for DIHM are approximately three times larger for Fresnel holograms and two times larger for Fourier holograms, compared to those of off-axis DHM. Assuming a fixed FOV in the reconstructed holograms, DIHM has a higher maximum spatial frequency than off-axis DHM. Therefore, the resolvable details of an object in reconstructed holograms from DIHM are finer than those from off-axis DHM. Additionally, high-frequency components in reconstructed holograms from off-axis DHM undergo partial filtration during the extraction process of real images. To determine the depth position of biological samples, an autofocusing function is usually utilized to search for an in-focus image with the highest sharpness or GRA among the reconstructed holograms. Given that the SBP of reconstructed holograms from DIHM is higher than that of off-axis DHM, adopting a DIHM system is advantageous for reconstructing edge structures with high-frequency components and accurately determining their 3D positional information.

The NA of off-axis DHM systems is also limited due to the requirement for a minimum recording distance to distinguish between real, twin, and zero-order images. To overcome this limitation in spatial resolution for off-axis DHM, it is common practice to employ an objective lens with a high magnification ratio and NA to acquire detailed 3D phase information of test samples. However, increasing the magnification ratio of the objective lens leads to a decrease in the FOV of off-axis DHM. On the other hand, DIHM systems typically utilize an objective lens with a relatively lower magnification ratio and NA for analyzing 3D locations of microscale particulates from the reconstructed holograms, while maintaining sufficient spatial resolution. Thus, 3D dynamics of biological samples moving in a wide FOV can be tracked more precisely compared to off-axis DHM. Additionally, due to its simpler optical configuration and shorter recording distance requirements, building up experimental setups for DIHM systems is easier. Therefore, the DIHM system is suitable for the 4D tracking of biological cells.

Advantages of DIHM over other microscopic imaging techniques

Various microscopic imaging techniques have been developed to visualize the 3D structures of biological samples with high lateral and axial resolutions. Fluorescence microscopy techniques, including confocal microscopy [20], two-photon microscopy [220], multi-photon fluorescence microscopy [221], and structured illumination microscopy [222], enable the reconstruction of 3D morphology by stacking fluorescent images at different focal planes or scanning fluorescence point-by-point across the samples. Light-sheet microscopy is utilized to scan a test sample by irradiating a light sheet at different depths and angles [223]. Spatial light interference microscopy is used to measure nanoscale phase information and dynamics of live cells over periods ranging from seconds to days [224]. Differential-interference-contrast microscopy collects a set of 2D images of a thick sample by moving it through the direction of focus to obtain its 3D image [225]. These advanced 3D imaging techniques effectively visualize the 3D structures and long-term variations of biological samples. However, real-time monitoring of rapidly changing cell dynamics is somewhat limited due to the shallow DOF and the long scanning time associated with these 3D imaging techniques.

Compared to other 3D imaging techniques, DIHM enables simultaneous measurement of multiple biological cells located at different depth-wise positions. It can analyze not only unicellular organisms but also multicellular organisms, including marine plankton [27, 49, 50, 226, 227], embryos [228], and stem cells [229, 230]. Additionally, DIHM allows visualization of transparent thin tissue structures, such as human breast carcinoma [231] and human hepatocellular carcinoma tissues [232]. Unlike the volumetric scanning process used in other techniques, DIHM records holographic images of moving cells using a high-speed camera with a high frame rate. This facilitates effective analysis of the 3D dynamics of biological cells with high temporal resolution.

The specialized nature of DIHM in analyzing the 3D dynamics of individual biological cells makes it a highly effective tool for studying multicellular interactions and collective cell migration behaviors. For example, DIHM has been employed to investigate cell-cell interactions between intestinal pathogenic bacteria E. sakazakii and the probiotic Lactobacillus rhamnosus [171]. Additionally, it has been utilized to monitor the morphology and migration behaviors of cancer cells in large-scale 3D matrix gels [233]. Consequently, DIHM systems can be effectively used for analyzing cell-cell interactions and cell migrations in various microenvironments that are challenging to investigate with other 3D imaging techniques.

Limitations of DIHM

Nevertheless, DIHM does have certain technical and experimental limitations. The lateral and axial resolutions are diffraction-limited due to the optical configuration of DIHM systems. The FOV is typically limited to hundreds of microns due to the restricted SBP of DIHM. Moreover, the shallow depth of focus prevents full reconstruction of the morphological structure of elongated objects. Overlapped holographic signals from highly concentrated particles reduce the accuracy of their 3D localization measurements. To avoid unexpected optical aberrations, it is crucial for the medium containing a test sample to have a uniform refractive index and for the beam paths within it to be free from impurities that induce unnecessary light scattering. Therefore, meticulous arrangement of the experimental setup and cautious execution of experiments are essential prerequisites for obtaining clear holographic signals from test samples.

Conclusions and perspectives

In summary, the DIHM technique holds great promise as a 3D imaging method suitable for label-free identification and tracking of various biological cells at the microscale level. It enables quantitative analysis of diverse 3D dynamics exhibited by biological cells, such as motility, migration, cell-surface interaction, and chemotactic behavior. By acquiring statistical information on their 3D location, orientation, and morphology over time with the aid of DIHM, it is possible to gain insights into these dynamic behaviors. Recent advancements in hologram handling techniques have significantly improved measurement accuracy in characterizing 3D dynamic behaviors of different species of cells under varying conditions such as diseases, external stimuli, and surrounding environments. Notably, rapid progress in AI technology has greatly enhanced the processing speed during reconstruction and autofocusing stages for precise 3D localization, while improving image quality through super-resolution algorithms and twin-image suppression methods. The reconstructed amplitude and phase maps derived from biological cells provide morphological and optical features that can be used for label-free detection, classification, and enumeration.

On the contrary, the majority of reconstruction and autofocusing methods that utilize AI techniques exhibit diminished generalization performance due to the limited availability of training datasets. Although AI techniques can effectively reduce the computational time required for reconstruction and autofocusing procedures, only a few studies have applied them to 3D PTV analysis of biological cells. Given the importance of demonstrating distinct differences in experimental samples and control groups in biological research, it is crucial to rapidly process numerous holographic images to obtain reliable statistical results. Conventional numerical reconstruction equations have been used for decades, as they are capable of reconstructing the amplitude and phase maps at different depth-wise locations from holographic signals of any unknown particles. If an innovative AI model with superior generalization performance were developed to replace conventional reconstruction methods, it would greatly accelerate the time-consuming digital image processing routines for cell tracking and facilitate related biological research.

The lateral and axial resolutions of DIHM are primarily determined by the NA of the DIHM setup and the wavelength of the light source, resulting in minor differences in resolutions among different DIHM systems. These slight variations depend on their specific optical configurations and numerical reconstruction methods. Rayleigh-Sommerfeld back-propagation and angular spectrum method would offer slightly more accurate holograms reconstructed from an original hologram recorded by DIHM systems using plane waves, as they do not require any approximation. Additionally, other numerical reconstruction equations using approximations can be utilized to reduce computational costs.

The main source of measurement errors in the 3D localization of biological cells is primarily associated with the autofocusing process. Although several autofocusing methods have demonstrated excellent performance for their own optical setup in DHM, it should be noted that the optimal focus function may depend on various factors, such as sample type, optical configuration, and post-image-processing method involving denoising and reconstruction algorithms. Therefore, selecting the most suitable autofocusing method for a specific experimental setup often requires trial and error. With the exception of highly concentrated particles, all PTV algorithms generally provide sufficient accuracy in the tracking process. It is recommended to refer to other studies that successfully analyzed similar biological cells to identify an optimal experimental setup and adapt it accordingly to suit specific experimental conditions. Several commercial DIHM platforms, including 4-Deep in-line holographic microscopes (NanoAndMore, USA), HO-DIHM-HT01 (Holmarc Opto-Mechatronics, India), and LISST-Holo2 (Sequoia Scientific, USA), are available for various applications.

By integrating the compact optical configurations of DIHM with state-of-the-art AI algorithms, it is possible to develop user-friendly devices for real-time and in situ analysis of biological cells. Portable smartphones or tablet computers equipped with DIHM and AI techniques can be effectively utilized for the facile diagnosis of hematologic diseases characterized by morphological disorders in human blood cells, as well as for continuous monitoring of hazardous microscale particulates, such as toxic contaminants, bacteria, and viruses in airborne or submersible environments. The ongoing advancements in DIHM and AI techniques hold great potential to significantly enhance the measurement performance related to translational and rotational behaviors, cell deformations, as well as cell-cell and cell-surface interactions. While previous studies on hemodynamics have primarily focused on the translational motions of erythrocytes, innovative AI-based DIHM techniques can now be employed to measure the dynamic behaviors of abnormal erythrocytes, thereby shedding light on their unknown impacts on human health. Furthermore, these advances enable simultaneous measurements of rapid variations in motility, orientation, and morphology of biological cells under various chemical, optical, and mechanical stimuli. Such progress offers high measurement accuracy along with significant throughput capacity and fast processing speed for quantitative analyses of biological cells. By fostering collaboration between engineers utilizing DIHM technologies and cell biologists providing expertise in cellular behaviors, a wide range of new research can be conducted to elucidate previously undisclosed dynamic characteristics of various biological cells, and will provide valuable information for potential applications in the diagnosis and treatment of associated diseases.

Availability of data and materials

Not applicable.

Abbreviations

2D:

Two-dimensional

3D:

Three-dimensional

4D:

Four-dimensional

AI:

Artificial intelligence

A. punctulata :

Arbacia punctulata

A. tumefaciens :

Agrobacterium tumefaciens

CNN:

Convolutional neural network

C. polykrikoides :

Cochlodinium polykrikoides

DHM:

Digital holographic microscopy

DIHM:

Digital in-line holographic microscopy

DOF:

Depth of field

D. primolecta :

Dunaliella primolecta

E. coli :

Escherichia coli

E. sakazakii :

Enterobacter sakazakii

FFT:

Fast Fourier transform

FOV:

Field of view

GRA:

Gradient

K. veneficum :

Karlodinium veneficum

LAP:

Laplacian

NA:

Numerical aperture

PDF:

Probability density function

PTV:

Particle tracking velocimetry

P. aeruginosa :

Pseudomonas aeruginosa

P. lima :

Prorocentrum lima

P. minimum :

Prorocentrum minimum

P. piscicida :

Pfiesteria piscicida

P. tricornutum :

Phaeodactylum tricornutum

P-value:

Probability value

SBP:

Space bandwidth product

SPEC:

Weighted spectral

SVM:

Support vector machine

S. putrefaciens :

Shewanella putrefaciens

TC:

Tamura coefficient

TTLL:

Tubulin-tyrosine ligase-like

Ttll3−/− Ttll8−/− mouse:

A double-knockout mouse for two initiating glycylases TTLL3 and TTLL8

VAR:

Variance

References

  1. Stephens DJ, Allan VJ. Light microscopy techniques for live cell imaging. Science. 2003;300(5616):82–6.

    Article  CAS  PubMed  Google Scholar 

  2. Ntziachristos V. Going deeper than microscopy: the optical imaging frontier in biology. Nat Methods. 2010;7(8):603–14.

    Article  CAS  PubMed  Google Scholar 

  3. Spiller DG, Wood CD, Rand DA, White MR. Measurement of single-cell dynamics. Nature. 2010;465(7299):736–45.

    Article  CAS  PubMed  Google Scholar 

  4. Cierpka C, Kähler CJ. Particle imaging techniques for volumetric three-component (3D3C) velocity measurements in microfluidics. J Vis. 2012;15:1–31.

    Article  CAS  Google Scholar 

  5. Morris JD, Payne CK. Microscopy and cell biology: new methods and new questions. Ann Rev Phys Chem. 2019;70:199–218.

    Article  CAS  Google Scholar 

  6. Gravesen P, Branebjerg J, Jensen OS. Microfluidics - a review. J Micromech Microeng. 1993;3(4):168.

    Article  CAS  Google Scholar 

  7. Stone HA, Stroock AD, Ajdari A. Engineering flows in small devices: microfluidics toward a lab-on-a-chip. Annu Rev Fluid Mech. 2004;36:381–411.

    Article  Google Scholar 

  8. Sackmann EK, Fulton AL, Beebe DJ. The present and future role of microfluidics in biomedical research. Nature. 2014;507(7491):181–9.

    Article  CAS  PubMed  Google Scholar 

  9. Samiei E, Tabrizian M, Hoorfar M. A review of digital microfluidics as portable platforms for lab-on a-chip applications. Lab Chip. 2016;16(13):2376–96.

    Article  CAS  PubMed  Google Scholar 

  10. Wang X, Hong XZ, Li YW, Li Y, Wang J, Chen P, et al. Microfluidics-based strategies for molecular diagnostics of infectious diseases. Mil Med Res. 2022;9(1):11.

    PubMed  PubMed Central  Google Scholar 

  11. Dupire J, Socol M, Viallat A. Full dynamics of a red blood cell in shear flow. Proc Natl Acad Sci U S A. 2012;109(51):20808–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Zeng NF, Ristenpart WD. Mechanical response of red blood cells entering a constriction. Biomicrofluidics. 2014;8(6):064123.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Lansche C, Dasanna AK, Quadt K, Fröhlich B, Missirlis D, Tétard M, et al. The sickle cell trait affects contact dynamics and endothelial cell activation in Plasmodium falciparum-infected erythrocytes. Commun Biol. 2018;1:211.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Atwell S, Badens C, Charrier A, Helfer E, Viallat A. Dynamics of individual red blood cells under shear flow: a way to discriminate deformability alterations. Front Physiol. 2022;12:775584.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Recktenwald SM, Graessel K, Maurer FM, John T, Gekle S, Wagner C. Red blood cell shape transitions and dynamics in time-dependent capillary flows. Biophys J. 2022;121(1):23–36.

    Article  CAS  PubMed  Google Scholar 

  16. Darrin M, Samudre A, Sahun M, Atwell S, Badens C, Charrier A, et al. Classification of red cell dynamics with convolutional and recurrent neural networks: a sickle cell disease case study. Sci Rep. 2023;13(1):745.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Egeblad M, Ewald AJ, Askautrud HA, Truitt ML, Welm BE, Bainbridge E, et al. Visualizing stromal cell dynamics in different tumor microenvironments by spinning disk confocal microscopy. Dis Models Mech. 2008;1(2–3):155–67.

    Article  Google Scholar 

  18. Renaud O, Herbomel P, Kissa K. Studying cell behavior in whole zebrafish embryos by confocal live imaging: application to hematopoietic stem cells. Nat Protoc. 2011;6(12):1897–904.

    Article  CAS  PubMed  Google Scholar 

  19. Carey SP, Kraning-Rush CM, Williams RM, Reinhart-King CA. Biophysical control of invasive tumor cell behavior by extracellular matrix microarchitecture. Biomaterials. 2012;33(16):4157–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Elliott AD. Confocal microscopy: principles and modern practices. Curr Protoc Cytom. 2020;92(1):e68.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Agero U, Monken C, Ropert C, Gazzinelli R, Mesquita O. Cell surface fluctuations studied with defocusing microscopy. Phys Rev E. 2003;67(5):051904.

    Article  CAS  Google Scholar 

  22. Etcheverry S, Gallardo MJ, Solano P, Suwalsky M, Mesquita ON, Saavedra C. Real-time study of shape and thermal fluctuations in the echinocyte transformation of human erythrocytes using defocusing microscopy. J Biomed Opt. 2012;17(10):106013.

    Article  PubMed  Google Scholar 

  23. Roma PM, Siman L, Hissa B, Agero U, Braga EM, Mesquita ON. Profiling of individual human red blood cells under osmotic stress using defocusing microscopy. J Biomed Opt. 2016;21(9):090505.

    Article  Google Scholar 

  24. Gabor D. A new microscopic principle. Nature. 1948;161(4098):777–8.

    Article  CAS  PubMed  Google Scholar 

  25. Murata S, Yasuda N. Potential of digital holography in particle measurement. Opt Laser Technol. 2000;32(7–8):567–74.

    Article  CAS  Google Scholar 

  26. Garcia-Sucerquia J, Xu W, Jericho SK, Klages P, Jericho MH, Kreuzer HJ. Digital in-line holographic microscopy. Appl Opt. 2006;45(5):836–50.

    Article  PubMed  Google Scholar 

  27. Sheng J, Malkiel E, Katz J. Digital holographic microscope for measuring three-dimensional particle distributions and motions. Appl Opt. 2006;45(16):3893–901.

    Article  PubMed  Google Scholar 

  28. Kim MK. Principles and techniques of digital holographic microscopy. SPIE Rev. 2010;1(1):018005.

    Google Scholar 

  29. Choi YS, Seo KW, Sohn MH, Lee SJ. Advances in digital holographic micro-PTV for analyzing microscale flows. Opt Lasers Eng. 2012;50(1):39–45.

    Article  Google Scholar 

  30. Yu X, Hong J, Liu C, Kim MK. Review of digital holographic microscopy for three-dimensional profiling and tracking. Opt Eng. 2014;53(11):112306.

    Article  Google Scholar 

  31. Memmolo P, Miccio L, Paturzo M, Di Caprio G, Coppola G, Netti PA, et al. Recent advances in holographic 3D particle tracking. Adv Opt Photonics. 2015;7(4):713–55.

    Article  Google Scholar 

  32. Wu Y, Ozcan A. Lensless digital holographic microscopy and its applications in biomedicine and environmental monitoring. Methods. 2018;136:4–16.

    Article  CAS  PubMed  Google Scholar 

  33. Xu W, Jericho M, Meinertzhagen I, Kreuzer H. Digital in-line holography of microspheres. Appl Opt. 2002;41(25):5367–75.

    Article  CAS  PubMed  Google Scholar 

  34. Sheng J, Malkiel E, Katz J. Using digital holographic microscopy for simultaneous measurements of 3D near wall velocity and wall shear stress in a turbulent boundary layer. Exp Fluids. 2008;45:1023–35.

    Article  Google Scholar 

  35. Choi YS, Lee SJ. Holographic analysis of three-dimensional inertial migration of spherical particles in micro-scale pipe flow. Microfluid Nanofluidics. 2010;9:819–29.

    Article  CAS  Google Scholar 

  36. Katz J, Sheng J. Applications of holography in fluid mechanics and particle dynamics. Annu Rev Fluid Mech. 2010;42:531–55.

    Article  Google Scholar 

  37. Choi YS, Seo KW, Lee SJ. Lateral and cross-lateral focusing of spherical particles in a square microchannel. Lab Chip. 2011;11(3):460–5.

    Article  CAS  PubMed  Google Scholar 

  38. Seo KW, Choi YS, Lee SJ. Dean-coupled inertial migration and transient focusing of particles in a curved microscale pipe flow. Exp Fluids. 2012;53(6):1867–77.

    Article  CAS  Google Scholar 

  39. Seo KW, Byeon HJ, Huh HK, Lee SJ. Particle migration and single-line particle focusing in microscale pipe flow of viscoelastic fluids. RSC Adv. 2014;4:3512–20.

    Article  CAS  Google Scholar 

  40. Lee SH, Roichman Y, Yi GR, Kim SH, Yang SM, van Blaaderen A, et al. Characterizing and tracking single colloidal particles with video holographic microscopy. Opt Express. 2007;15(26):18275–82.

    Article  PubMed  Google Scholar 

  41. Fung J, Martin KE, Perry RW, Kaz DM, McGorty R, Manoharan VN. Measuring translational, rotational, and vibrational dynamics in colloids with digital holographic microscopy. Opt Express. 2011;19(9):8051–65.

    Article  PubMed  Google Scholar 

  42. Verrier N, Fournier C, Fournel T. 3D tracking the Brownian motion of colloidal particles using digital holographic microscopy and joint reconstruction. Appl Opt. 2015;54(16):4996–5002.

    Article  PubMed  Google Scholar 

  43. Ling H, Srinivasan S, Golovin K, McKinley GH, Tuteja A, Katz J. High-resolution velocity measurement in the inner part of turbulent boundary layers over super-hydrophobic surfaces. J Fluid Mech. 2016;801:670–703.

    Article  CAS  Google Scholar 

  44. Wang L, Wu Y, Wu X, Cen K. Measurement of dynamics of laser-induced cavitation around nanoparticle with high-speed digital holographic microscopy. Exp Therm Fluid Sci. 2021;121:110266.

    Article  CAS  Google Scholar 

  45. Go T, Kim J, Lee SJ. Three-dimensional volumetric monitoring of settling particulate matters on a leaf using digital in-line holographic microscopy. J Hazard Mater. 2021;404(Pt A):124116.

    Article  CAS  PubMed  Google Scholar 

  46. Kim J, Kim J, Kim Y, Go T, Lee SJ. Accelerated settling velocity of airborne particulate matter on hairy plant leaves. J Environ Manage. 2023;332:117313.

    Article  CAS  PubMed  Google Scholar 

  47. Xu W, Jericho M, Meinertzhagen I, Kreuzer H. Digital in-line holography for biological applications. Proc Natl Acad Sci U S A. 2001;98(20):11301–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Jericho S, Garcia-Sucerquia J, Xu W, Jericho M, Kreuzer H. Submersible digital in-line holographic microscope. Rev Sci Instrum. 2006;77(4):043706.

    Article  Google Scholar 

  49. Rotermund L, Samson J, Kreuzer H. A submersible holographic microscope for 4D in-situ studies of micro-organisms in the ocean with intensity and quantitative phase imaging. J Marine Sci Res Dev. 2016;6(1):1000181.

    Google Scholar 

  50. Liu Z, Takahashi T, Lindsay D, Thevar T, Sangekar M, Watanabe HK, et al. Digital in-line holography for large-volume analysis of vertical motion of microscale marine plankton and other particles. IEEE J Ocean Eng. 2021;46(4):1248–60.

    Article  Google Scholar 

  51. Repetto L, Piano E, Pontiggia C. Lensless digital holographic microscope with light-emitting diode illumination. Opt Lett. 2004;29(10):1132–4.

    Article  CAS  PubMed  Google Scholar 

  52. Mariën J, Stahl R, Lambrechts A, van Hoof C, Yurt A. Color lens-free imaging using multi-wavelength illumination based phase retrieval. Opt Express. 2020;28(22):33002–18.

    Article  PubMed  Google Scholar 

  53. Xiong Z, Potter CJ, McLeod E. High-speed lens-free holographic sensing of protein molecules using quantitative agglutination assays. ACS Sensors. 2021;6(3):1208–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Liu J-P, Tahara T, Hayasaki Y, Poon TC. Incoherent digital holography: a review. Appl Sci. 2018;8(1):143.

    Article  Google Scholar 

  55. Rosen J, Vijayakumar A, Kumar M, Rai MR, Kelner R, Kashter Y, et al. Recent advances in self-interference incoherent digital holography. Adv Opt Photonics. 2019;11(1):1–66.

    Article  Google Scholar 

  56. Tahara T, Zhang Y, Rosen J, Anand V, Cao L, Wu J, et al. Roadmap of incoherent digital holography. Appl Phys B. 2022;128:193.

    Article  CAS  Google Scholar 

  57. Tahara T. Polarization-filterless polarization-sensitive polarization-multiplexed phase-shifting incoherent digital holography (P4IDH). Opt Lett. 2023;48(15):3881–4.

    Article  CAS  PubMed  Google Scholar 

  58. Chang M, Hu CP, Lam P, Wyant JC. High precision deformation measurement by digital phase shifting holographic interferometry. Appl Opt. 1985;24(22):3780–3.

    Article  CAS  PubMed  Google Scholar 

  59. Awatsuji Y, Tahara T, Kaneko A, Koyama T, Nishio K, Ura S, et al. Parallel two-step phase-shifting digital holography. Appl Opt. 2008;47(19):D183–9.

    Article  PubMed  Google Scholar 

  60. Tahara T, Awatsuji Y, Shimozato Y, Kakue T, Nishio K, Ura S, et al. Single-shot polarization-imaging digital holography based on simultaneous phase-shifting interferometry. Opt Lett. 2011;36(16):3254–6.

    Article  PubMed  Google Scholar 

  61. Sanz M, Picazo-Bueno JA, García J, Micó V. Improved quantitative phase imaging in lensless microscopy by single-shot multi-wavelength illumination using a fast convergence algorithm. Opt Express. 2015;23(16):21352–65.

    Article  CAS  PubMed  Google Scholar 

  62. Farthing NE, Findlay RC, Jikeli JF, Walrad PB, Bees MA, Wilson LG. Simultaneous two-color imaging in digital holographic microscopy. Opt Express. 2017;25(23):28489–500.

    Article  PubMed  Google Scholar 

  63. Min J, Zhou M, Yuan X, Wen K, Yu X, Peng T, et al. Optical thickness measurement with single-shot dual-wavelength in-line digital holography. Opt Lett. 2018;43(18):4469–72.

    Article  CAS  PubMed  Google Scholar 

  64. Zhang H, Stangner T, Wiklund K, Andersson M. Object plane detection and phase retrieval from single-shot holograms using multi-wavelength in-line holography. Appl Opt. 2018;57(33):9855–62.

    Article  PubMed  Google Scholar 

  65. Lee SH, Grier DG. Holographic microscopy of holographically trapped three-dimensional structures. Opt Express. 2007;15(4):1505–12.

    Article  PubMed  Google Scholar 

  66. Cheong FC, Xiao K, Grier DG. Characterizing individual milk fat globules with holographic video microscopy. J Dairy Sci. 2009;92(1):95–9.

    Article  CAS  PubMed  Google Scholar 

  67. Cheong FC, Krishnatreya BJ, Grier DG. Strategies for three-dimensional particle tracking with holographic video microscopy. Opt Express. 2010;18(13):13563–73.

    Article  PubMed  Google Scholar 

  68. Cheong FC, Grier DG. Rotational and translational diffusion of copper oxide nanorods measured with holographic video microscopy. Opt Express. 2010;18(7):6555–62.

    Article  CAS  PubMed  Google Scholar 

  69. Fugal JP, Schulz TJ, Shaw RA. Practical methods for automated reconstruction and characterization of particles in digital in-line holograms. Meas Sci Technol. 2009;20(7):075501.

    Article  Google Scholar 

  70. Pedrini G, Schedin S, Tiziani HJ. Spatial filtering in digital holographic microscopy. J Mod Opt. 2000;47(8):1447–54.

    Article  Google Scholar 

  71. Malkiel E, Abras JN, Katz J. Automated scanning and measurements of particle distributions within a holographic reconstructed volume. Meas Sci Technol. 2004;15(4):601.

    Article  CAS  Google Scholar 

  72. Lee SJ, Seo KW, Choi YS, Sohn MH. Three-dimensional motion measurements of free-swimming microorganisms using digital holographic microscopy. Meas Sci Technol. 2011;22:064004.

    Article  Google Scholar 

  73. Singh DK, Panigrahi P. Improved digital holographic reconstruction algorithm for depth error reduction and elimination of out-of-focus particles. Opt Express. 2010;18(3):2426–48.

    Article  PubMed  Google Scholar 

  74. Garcia-Sucerquia J, Ramírez JAH, Prieto DV. Reduction of speckle noise in digital holography by using digital image processing. Optik. 2005;116(1):44–8.

    Article  Google Scholar 

  75. Yang Y, Kang BS, Choo YJ. Application of the correlation coefficient method for determination of the focal plane to digital particle holography. Appl Opt. 2008;47(6):817–24.

    Article  PubMed  Google Scholar 

  76. Kukrer O, Hocanin A. Frequency-response-shaped LMS adaptive filter. Digit Signal Process. 2006;16(6):855–69.

    Article  Google Scholar 

  77. Zajtsev AK, Lin SH, Hsu KY. Sidelobe suppression of spectral response in holographic optical filter. Opt Commun. 2001;190(1–6):103–8.

    Article  CAS  Google Scholar 

  78. Sharma A, Sheoran G, Jaffery Z. Improvement of signal-to-noise ratio in digital holography using wavelet transform. Opt Lasers Eng. 2008;46(1):42–7.

    Article  Google Scholar 

  79. Uzan A, Rivenson Y, Stern A. Speckle denoising in digital holography by nonlocal means filtering. Appl Opt. 2013;52(1):A195–200.

    Article  PubMed  Google Scholar 

  80. Molaei M, Sheng J. Imaging bacterial 3D motion using digital in-line holographic microscopy and correlation-based de-noising algorithm. Opt Express. 2014;22(26):32119–37.

    Article  PubMed  PubMed Central  Google Scholar 

  81. Jeon W, Jeong W, Son K, Yang H. Speckle noise reduction for digital holographic images using multi-scale convolutional neural networks. Opt Lett. 2018;43(17):4240–3.

    Article  PubMed  Google Scholar 

  82. Bai C, Peng T, Min J, Li R, Zhou Y, Yao B. Dual-wavelength in-line digital holography with untrained deep neural networks. Photonics Res. 2021;9(12):2501–10.

    Article  Google Scholar 

  83. Chen L, Chen X, Cui H, Long Y, Wu J. Image enhancement in lensless inline holographic microscope by inter-modality learning with denoising convolutional neural network. Opt Commun. 2021;484:126682.

    Article  CAS  Google Scholar 

  84. Bishara W, Su TW, Coskun AF, Ozcan A. Lensfree on-chip microscopy over a wide field-of-view using pixel super-resolution. Opt Express. 2010;18(11):11181–91.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Byeon H, Go T, Lee SJ. Deep learning-based digital in-line holographic microscopy for high resolution with extended field of view. Opt Laser Technol. 2019;113:77–86.

    Article  CAS  Google Scholar 

  86. Luo Z, Yurt A, Stahl R, Lambrechts A, Reumers V, Braeken D, et al. Pixel super-resolution for lens-free holographic microscopy using deep learning neural networks. Opt Express. 2019;27(10):13581–95.

    Article  CAS  PubMed  Google Scholar 

  87. Lee H, Kim J, Kim J, Jeon P, Lee SA, Kim D. Noniterative sub-pixel shifting super-resolution lensless digital holography. Opt Express. 2021;29(19):29996–30006.

    Article  CAS  PubMed  Google Scholar 

  88. Potter CJ, Hu Y, Xiong Z, Wang J, McLeod E. Point-of-care SARS-CoV-2 sensing using lens-free imaging and a deep learning-assisted quantitative agglutination assay. Lab Chip. 2022;22(19):3744–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Goodman JW. Introduction to Fourier optics. Colorado: Roberts and Company Publishers; 2005.

    Google Scholar 

  90. Born M, Wolf E. Principles of optics: electromagnetic theory of propagation, interference and diffraction of light. Amsterdam: Elsevier; 2013.

    Google Scholar 

  91. Barton JJ. Photoelectron holography. Phys Rev Lett. 1988;61(12):1356–9.

    Article  CAS  PubMed  Google Scholar 

  92. Kreuzer H, Nakamura K, Wierzbicki A, Fink H, Schmid H. Theory of the point source electron microscope. Ultramicroscopy. 1992;45(3–4):381–403.

    Article  Google Scholar 

  93. Kreuzer H. Low energy electron point source microscopy. Micron. 1995;26(6):503–9.

    Article  Google Scholar 

  94. Delen N, Hooker B. Free-space beam propagation between arbitrarily oriented planes based on full diffraction theory: a fast Fourier transform approach. JOSA A. 1998;15(4):857–67.

    Article  Google Scholar 

  95. Veerman JA, Rusch JJ, Urbach HP. Calculation of the Rayleigh-Sommerfeld diffraction integral by exact integration of the fast oscillating factor. JOSA A. 2005;22(4):636–46.

    Article  PubMed  Google Scholar 

  96. Shen F, Wang A. Fast-Fourier-transform based numerical integration method for the Rayleigh-Sommerfeld diffraction formula. Appl Opt. 2006;45(6):1102–10.

    Article  PubMed  Google Scholar 

  97. Wilson L, Zhang R. 3D Localization of weak scatterers in digital holographic microscopy using Rayleigh-Sommerfeld back-propagation. Opt Express. 2012;20(15):16735–44.

    Article  Google Scholar 

  98. Sheng J, Malkiel E, Katz J, Adolf J, Belas R, Place AR. Digital holographic microscopy reveals prey-induced changes in swimming behavior of predatory dinoflagellates. Proc Natl Acad Sci U S A. 2007;104(44):17512–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Ratcliffe JA. Some aspects of diffraction theory and their application to the ionosphere. Rep Prog Phys. 1956;19:188.

    Article  Google Scholar 

  100. Koren G, Polack F, Joyeux D. Iterative algorithms for twin-image elimination in in-line holography using finite-support constraints. JOSA A. 1993;10(3):423–33.

    Article  Google Scholar 

  101. Latychevskaia T, Fink HW. Solution to the twin image problem in holography. Phys Rev Lett. 2007;98(23):233901.

    Article  PubMed  Google Scholar 

  102. Ling H, Katz J. Separating twin images and locating the center of a microparticle in dense suspensions using correlations among reconstructed fields of two parallel holograms. Appl Opt. 2014;53(27):G1–11.

    Article  PubMed  Google Scholar 

  103. Oe K, Nomura T. Twin-image reduction method using a diffuser for phase imaging in-line digital holography. Appl Opt. 2018;57(20):5652–6.

    Article  PubMed  Google Scholar 

  104. Rivenson Y, Zhang Y, Günaydın H, Teng D, Ozcan A. Phase recovery and holographic image reconstruction using deep learning in neural networks. Light Sci Appl. 2018;7:17141.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Latychevskaia T. Iterative phase retrieval for digital holography: tutorial. JOSA A. 2019;36(12):D31–40.

    Article  CAS  PubMed  Google Scholar 

  106. Shangraw M, Ling H. Separating twin images in digital holographic microscopy using weak scatterers. Appl Opt. 2021;60(3):626–34.

    Article  PubMed  Google Scholar 

  107. Langehanenberg P, Kemper B, Dirksen D, von Bally G. Autofocusing in digital holographic phase contrast microscopy on pure phase objects for live cell imaging. Appl Opt. 2008;47(19):D176–82.

    Article  PubMed  Google Scholar 

  108. Memmolo P, Distante C, Paturzo M, Finizio A, Ferraro P, Javidi B. Automatic focusing in digital holography and its application to stretched holograms. Opt Lett. 2011;36(10):1945–7.

    Article  CAS  PubMed  Google Scholar 

  109. Gibson T, Bedrossian M, Serabyn E, Lindensmith C, Nadeau JL. Using the Gouy phase anomaly to localize and track bacteria in digital holographic microscopy 4D images. JOSA A. 2021;38(2):A11–8.

    Article  CAS  PubMed  Google Scholar 

  110. Memmolo P, Paturzo M, Javidi B, Netti PA, Ferraro P. Refocusing criterion via sparsity measurements in digital holography. Opt Lett. 2014;39(16):4719–22.

    Article  PubMed  Google Scholar 

  111. Yeo T, Ong S, Sinniah R. Autofocusing for tissue microscopy. Image Vis Comput. 1993;11(10):629–39.

    Article  Google Scholar 

  112. Brenner JF, Dew BS, Horton JB, King T, Neurath PW, Selles WD. An automated microscope for cytologic research a preliminary evaluation. J Histochem Cytochem. 1976;24(1):100–11.

    Article  CAS  PubMed  Google Scholar 

  113. Trusiak M, Picazo-Bueno JA, Zdankowski P, Micó V. DarkFocus: numerical autofocusing in digital in-line holographic microscopy using variance of computational dark-field gradient. Opt Lasers Eng. 2020;134:106195.

    Article  Google Scholar 

  114. Li W, Loomis NC, Hu Q, Davis CS. Focus detection from digital in-line holograms based on spectral L1 norms. JOSA A. 2007;24(10):3054–62.

    Article  PubMed  Google Scholar 

  115. Kumar SS, Sun Y, Zou S, Hong J. 3D holographic observatory for long-term monitoring of complex behaviors in drosophila. Sci Rep. 2016;6:33001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Ren Z, Xu Z, Lam EY. Learning-based nonparametric autofocusing for digital holography. Optica. 2018;5(4):337–44.

    Article  Google Scholar 

  117. Wu Y, Rivenson Y, Zhang Y, Wei Z, Günaydin H, Lin X, et al. Extended depth-of-field in holographic imaging using deep-learning-based autofocusing and phase recovery. Optica. 2018;5(6):704–10.

    Article  Google Scholar 

  118. Lee SJ, Yoon GY, Go T. Deep learning-based accurate and rapid tracking of 3D positional information of microparticles using digital holographic microscopy. Exp Fluids. 2019;60:170.

    Article  CAS  Google Scholar 

  119. Pitkäaho T, Manninen A, Naughton TJ. Focus prediction in digital holographic microscopy using deep convolutional neural networks. Appl Opt. 2019;58(5):A202–8.

    Article  PubMed  Google Scholar 

  120. Montoya M, Lopera MJ, Gómez-Ramírez A, Buitrago-Duque C, Pabón-Vidal A, Herrera-Ramirez J, et al. FocusNET: an autofocusing learning-based model for digital lensless holographic microscopy. Opt Lasers Eng. 2023;165:107546.

    Article  Google Scholar 

  121. Baek S, Lee S. A new two-frame particle tracking algorithm using match probability. Exp Fluids. 1996;22:23–32.

    Article  CAS  Google Scholar 

  122. Crocker JC, Grier DG. Methods of digital video microscopy for colloidal studies. J Colloid Interface Sci. 1996;179(1):298–310.

    Article  CAS  Google Scholar 

  123. Allan DB, Caswell T, Keim N, van der Wel C, Verweij R. Soft-matter/trackpy: v0.6.1. Zenodo; 2023. https://zenodo.org/records/7670439.

  124. Hassan Y, Canaan R. Full-field bubbly flow velocity measurements using a multiframe particle tracking technique. Exp Fluids. 1991;12:49–60.

    Article  CAS  Google Scholar 

  125. Malik N, Dracos T, Papantoniou D. Particle tracking velocimetry in three-dimensional flows. Exp Fluids. 1993;15:279–94.

    Article  CAS  Google Scholar 

  126. Ouellette NT, Xu H, Bodenschatz E. A quantitative study of three-dimensional Lagrangian particle tracking algorithms. Exp Fluids. 2006;40:301–13.

    Article  Google Scholar 

  127. Li D, Zhang Y, Sun Y, Yan W. A multi-frame particle tracking algorithm robust against input noise. Meas Sci Technol. 2008;19(10):105401.

    Article  Google Scholar 

  128. Cierpka C, Lütke B, Kähler CJ. Higher order multi-frame particle tracking velocimetry. Exp Fluids. 2013;54:1533.

    Article  Google Scholar 

  129. Labonté G. Neural network reconstruction of fluid flows from tracer-particle displacements. Exp Fluids. 2001;30:399–409.

    Article  Google Scholar 

  130. Mallery K, Shao S, Hong J. Dense particle tracking using a learned predictive model. Exp Fluids. 2020;61:223.

    Article  Google Scholar 

  131. Dixon L, Cheong FC, Grier DG. Holographic deconvolution microscopy for high-resolution particle tracking. Opt Express. 2011;19(17):16410–7.

    Article  CAS  PubMed  Google Scholar 

  132. Latychevskaia T, Fink HW. Holographic time-resolved particle tracking by means of three-dimensional volumetric deconvolution. Opt Express. 2014;22(17):20994–1003.

    Article  PubMed  Google Scholar 

  133. Toloui M, Hong J. High fidelity digital inline holographic method for 3D flow measurements. Opt Express. 2015;23(21):27159–73.

    Article  CAS  PubMed  Google Scholar 

  134. Mallery K, Hong J. Regularized inverse holographic volume reconstruction for 3D particle tracking. Opt Express. 2019;27(13):18069–84.

    Article  CAS  PubMed  Google Scholar 

  135. Chen N, Wang C, Heidrich W. Snapshot space-time holographic 3D particle tracking velocimetry. Laser Photonics Rev. 2021;15(8):2100008.

    Article  Google Scholar 

  136. Sun B, Ahmed A, Atkinson C, Soria J. A novel 4D digital holographic PIV/PTV (4D-DHPIV/PTV) methodology using iterative predictive inverse reconstruction. Meas Sci Technol. 2020;31(10):104002.

    Article  CAS  Google Scholar 

  137. Shao S, Mallery K, Kumar SS, Hong J. Machine learning holography for 3D particle field imaging. Opt Express. 2020;28(3):2987–99.

    Article  PubMed  Google Scholar 

  138. Wang K, Dou J, Kemao Q, Di J, Zhao J. Y-Net: a one-to-two deep learning framework for digital holographic reconstruction. Opt Lett. 2019;44(19):4765–8.

    Article  PubMed  Google Scholar 

  139. Yin D, Gu Z, Zhang Y, Gu F, Nie S, Ma J, et al. Digital holographic reconstruction based on deep learning framework with unpaired data. IEEE Photonics J. 2019;12(2):3900312.

    Google Scholar 

  140. Jaferzadeh K, Fevens T. HoloPhaseNet: fully automated deep-learning-based hologram reconstruction using a conditional generative adversarial model. Biomed Opt Express. 2022;13(7):4032–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Kiriy SA, Rymov DA, Svistunov AS, Shifrina AV, Starikov RS, Cheremkhin PA. Generative adversarial neural network for 3D-hologram reconstruction. Laser Phys Lett. 2024;21(4):045201.

    Article  Google Scholar 

  142. Chen H, Huang L, Liu T, Ozcan A. Fourier imager network (FIN): a deep neural network for hologram reconstruction with superior external generalization. Light Sci Appl. 2022;11(1):254.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Huang L, Chen H, Liu T, Ozcan A. Self-supervised learning of hologram reconstruction using physics consistency. Nat Mach Intell. 2023;5:895–907.

    Article  Google Scholar 

  144. Sun H, Song B, Dong H, Reid B, Player MA, Watson J, et al. Visualization of fast-moving cells in vivo using digital holographic video microscopy. J Biomed Opt. 2008;13(1):014007.

    Article  PubMed  Google Scholar 

  145. Nette F, Guerra de Souza AC, Laskay T, Ohms M, Dömer D, Drömann D, et al. Method for simultaneous tracking of thousands of unlabeled cells within a transparent 3D matrix. PLoS One. 2022;17(6):e0270456.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Choi YS, Lee SJ. Three-dimensional volumetric measurement of red blood cell motion using digital holographic microscopy. Appl Opt. 2009;48(16):2983–90.

    Article  PubMed  Google Scholar 

  147. Choi YS, Lee SJ. Inertial migration of erythrocytes in low-viscosity and high-shear rate microtube flows: aplication simple digital in-line holographic microscopy. J Biomech. 2012;45(15):2706–9.

    Article  PubMed  Google Scholar 

  148. Seo KW, Ha YR, Lee SJ. Vertical focusing and cell ordering in a microchannel via viscoelasticity: applications for cell monitoring using a digital holographic microscopy. Appl Phys Lett. 2014;104(21):213702.

    Article  Google Scholar 

  149. Go T, Byeon H, Lee SJ. Focusing and alignment of erythrocytes in a viscoelastic medium. Sci Rep. 2017;7:41162.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Byeon H, Go T, Lee SJ. Digital stereo-holographic microscopy for studying three-dimensional particle dynamics. Opt Lasers Eng. 2018;105:6–13.

    Article  Google Scholar 

  151. Kim Y, Kim J, Seo E, Lee SJ. AI-based analysis of 3D position and orientation of red blood cells using a digital in-line holographic microscopy. Biosens Bioelectron. 2023;229:115232.

    Article  CAS  PubMed  Google Scholar 

  152. Su TW, Xue L, Ozcan A. High-throughput lensfree 3D tracking of human sperms reveals rare statistics of helical trajectories. Proc Natl Acad Sci U S A. 2012;109(40):16018–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Picazo-Bueno JA, Trindade K, Sanz M, Micó V. Design, calibration, and application of a robust, cost-effective, and high-resolution lensless holographic microscope. Sensors. 2022;22(2):553.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Rogalski M, Picazo-Bueno JA, Winnik J, Zdańkowski P, Micó V, Trusiak M. Accurate automatic object 4D tracking in digital in-line holographic microscopy based on computationally rendered dark fields. Sci Rep. 2022;12:12909.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Jikeli JF, Alvarez L, Friedrich BM, Wilson LG, Pascal R, Colin R, et al. Sperm navigation along helical paths in 3D chemoattractant landscapes. Nat Commun. 2015;6:7985.

    Article  CAS  PubMed  Google Scholar 

  156. Su TW, Choi I, Feng J, Huang K, Ozcan A. High-throughput analysis of horse sperms’ 3D swimming patterns using computational on-chip imaging. Anim Reprod Sci. 2016;169:45–55.

    Article  PubMed  Google Scholar 

  157. Gadadhar S, Alvarez Viar G, Hansen JN, Gong A, Kostarev A, Ialy-Radio C, et al. Tubulin glycylation controls axonemal dynein activity, flagellar beat, and male fertility. Science. 2021;371(6525):eabd4914.

    Article  CAS  PubMed  Google Scholar 

  158. Vater SM, Weiße S, Maleschlijski S, Lotz C, Koschitzki F, Schwartz T, et al. Swimming behavior of Pseudomonas aeruginosa studied by holographic 3D tracking. PLoS One. 2014;9(1):e87765.

    Article  PubMed  PubMed Central  Google Scholar 

  159. Cheong FC, Wong CC, Gao Y, Nai MH, Cui Y, Park S, et al. Rapid, high-throughput tracking of bacterial motility in 3D via phase-contrast holographic video microscopy. Biophys J. 2015;108(5):1248–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Tai CW, Ahmadzadegan A, Ardekani A, Narsimhan V. A forward reconstruction, holographic method to overcome the lens effect during 3D detection of semi-transparent, non-spherical particles. Soft Matter. 2023;19(1):115–27.

    Article  CAS  Google Scholar 

  161. Wang A, Garmann RF, Manoharan VN. Tracking E. coli runs and tumbles with scattering solutions and digital holographic microscopy. Opt Express. 2016;24(21):23719–25.

    Article  PubMed  Google Scholar 

  162. Kühn MJ, Schmidt FK, Farthing NE, Rossmann FM, Helm B, Wilson LG, et al. Spatial arrangement of several flagellins within bacterial flagella improves motility in different environments. Nat Commun. 2018;9(1):5369.

    Article  PubMed  PubMed Central  Google Scholar 

  163. Molaei M, Barry M, Stocker R, Sheng J. Failed escape: solid surfaces prevent tumbling of Escherichia coli. Phys Rev Lett. 2014;113(6):068103.

    Article  PubMed  Google Scholar 

  164. Molaei M, Sheng J. Succeed escape: flow shear promotes tumbling of Escherichia coli near a solid surface. Sci Rep. 2016;6:35290.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Qi M, Gong X, Wu B, Zhang G. Landing dynamics of swimming bacteria on a polymeric surface: effect of surface properties. Langmuir. 2017;33(14):3525–33.

    Article  CAS  PubMed  Google Scholar 

  166. Bianchi S, Saglimbeni F, Di Leonardo R. Holographic imaging reveals the mechanism of wall entrapment in swimming bacteria. Phys Rev X. 2017;7(1):011010.

    Google Scholar 

  167. Qi M, Song Q, Zhao J, Ma C, Zhang G, Gong X. Three-dimensional bacterial behavior near dynamic surfaces formed by degradable polymers. Langmuir. 2017;33(45):13098–104.

    Article  CAS  PubMed  Google Scholar 

  168. Peng Q, Zhou X, Wang Z, Xie Q, Ma C, Zhang G, et al. Three-dimensional bacterial motions near a surface investigated by digital holographic microscopy: effect of surface stiffness. Langmuir. 2019;35(37):12257–63.

    Article  CAS  PubMed  Google Scholar 

  169. Hook AL, Flewellen JL, Dubern JF, Carabelli AM, Zaid IM, Berry RM, et al. Simultaneous tracking of Pseudomonas aeruginosa motility in liquid and at the solid-liquid interface reveals differential roles for the flagellar stators. mSystems. 2019;4(5):e00390–e419.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Elius M, Boyle K, Chang WS, Moisander PH, Ling H. Comparison of three-dimensional motion of bacteria with and without wall accumulation. Phys Rev E. 2023;108(1):014409.

    Article  CAS  PubMed  Google Scholar 

  171. He X, Zhang W, Feng P, Mai Z, Gong X, Zhang G. Role of surface coverage of sessile probiotics in their interplay with pathogen bacteria investigated by digital holographic microscopy. Langmuir. 2023;39(48):17308–17.

    Article  CAS  PubMed  Google Scholar 

  172. Sohn MH, Seo KW, Choi YS, Lee SJ, Kang YS, Kang YS. Determination of the swimming trajectory and speed of chain-forming dinoflagellate Cochlodinium polykrikoides with digital holographic particle tracking velocimetry. Mar Biol. 2011;158:561–70.

    Article  Google Scholar 

  173. Lee SJ, Go T, Byeon H. Three-dimensional swimming motility of microorganism in the near-wall region. Exp Fluids. 2016;57:26.

    Article  Google Scholar 

  174. Lewis NI, Xu W, Jericho SK, Kreuzer HJ, Jericho MH, Cembella AD. Swimming speed of three species of Alexandrium (Dinophyceae) as determined by digital in-line holography. Phycologia. 2006;45(1):61–70.

    Article  Google Scholar 

  175. Sheng J, Malkiel E, Katz J, Adolf JE, Place AR. A dinoflagellate exploits toxins to immobilize prey prior to ingestion. Proc Natl Acad Sci U S A. 2010;107(5):2082–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Sohn MH, Lim S, Seo KW, Lee SJ. Effect of ambient medium viscosity on the motility and flagella motion of Prorocentrum minimum (Dinophyceae). J Plankton Res. 2013;35(6):1294–304.

    Article  Google Scholar 

  177. Dharmawan AB, Mariana S, Scholz G, Hörmann P, Schulze T, Triyana K, et al. Nonmechanical parfocal and autofocus features based on wave propagation distribution in lensfree holographic microscopy. Sci Rep. 2021;11:3213.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Xu W, Jericho M, Kreuzer H, Meinertzhagen I. Tracking particles in four dimensions with in-line holographic microscopy. Opt Lett. 2003;28(3):164–6.

    Article  CAS  PubMed  Google Scholar 

  179. Jericho S, Klages P, Nadeau J, Dumas E, Jericho M, Kreuzer H. In-line digital holographic microscopy for terrestrial and exobiological research. Planet Space Sci. 2010;58(4):701–5.

    Article  Google Scholar 

  180. Lee SJ, Byeon HJ, Seo KW. Inertial migration of spherical elastic phytoplankton in pipe flow. Exp Fluids. 2014;55:1742.

    Article  Google Scholar 

  181. Chengala A, Hondzo M, Sheng J. Microalga propels along vorticity direction in a shear flow. Phys Rev E. 2013;87(5):052704.

    Article  Google Scholar 

  182. You J, Mallery K, Mashek DG, Sanders M, Hong J, Hondzo M. Microalgal swimming signatures and neutral lipids production across growth phases. Biotechnol Bioeng. 2020;117(4):970–80.

    Article  CAS  PubMed  Google Scholar 

  183. Heydt M, Rosenhahn A, Grunze M, Pettitt M, Callow M, Callow J. Digital in-line holography as a three-dimensional tool to study motile marine organisms during their exploration of surfaces. J Adhes. 2007;83(5):417–30.

    Article  CAS  Google Scholar 

  184. Heydt M, Divós P, Grunze M, Rosenhahn A. Analysis of holographic microscopy data to quantitatively investigate three-dimensional settlement dynamics of algal zoospores in the vicinity of surfaces. Eur Phys J E. 2009;30:141–8.

    Article  CAS  PubMed  Google Scholar 

  185. Heydt M, Pettitt M, Cao X, Callow M, Callow J, Grunze M, et al. Settlement behavior of zoospores of Ulva linza during surface selection studied by digital holographic microscopy. Biointerphases. 2012;7:33.

    Article  CAS  PubMed  Google Scholar 

  186. Weiße S, Heddergott N, Heydt M, Pflästerer D, Maier T, Haraszti T, et al. A quantitative 3D motility analysis of Trypanosoma brucei by use of digital in-line holographic microscopy. PLoS One. 2012;7(5):e37296.

    Article  PubMed  PubMed Central  Google Scholar 

  187. Thornton KL, Butler JK, Davis SJ, Baxter BK, Wilson LG. Haloarchaea swim slowly for optimal chemotactic efficiency in low nutrient environments. Nat Commun. 2020;11:4453.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Findlay RC, Osman M, Spence KA, Kaye PM, Walrad PB, Wilson LG. High-speed, three-dimensional imaging reveals chemotactic behaviour specific to human-infective Leishmania parasites. Elife. 2021;10:e65051.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Pal D, Amyot M, Liang C, Ariya PA. Real-time 4D tracking of airborne virus-laden droplets and aerosols. Commun Eng. 2023;2:41.

    Article  PubMed Central  Google Scholar 

  190. Seo S, Su TW, Tseng DK, Erlinger A, Ozcan A. Lensfree holographic imaging for on-chip cytometry and diagnostics. Lab Chip. 2009;9:777–87.

    Article  CAS  PubMed  Google Scholar 

  191. Seo S, Isikman SO, Sencan I, Mudanyali O, Su TW, Bishara W, et al. High-throughput lens-free blood analysis on a chip. Anal Chem. 2010;82(11):4621–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Vercruysse D, Dusa A, Stahl R, Vanmeerbeeck G, de Wijs K, Liu C, et al. Three-part differential of unlabeled leukocytes with a compact lens-free imaging flow cytometer. Lab Chip. 2015;15:1123–32.

    Article  CAS  PubMed  Google Scholar 

  193. Park JH, Go T, Lee SJ. Label-free sensing and classification of old stored blood. Ann Biomed Eng. 2017;45(11):2563–73.

    Article  PubMed  Google Scholar 

  194. Singh DK, Ahrens CC, Li W, Vanapalli SA. Label-free fingerprinting of tumor cells in bulk flow using inline digital holographic microscopy. Biomed Opt Express. 2017;8(2):536–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Buzalewicz I, Kujawińska M, Krauze W, Podbielska H. Novel perspectives on the characterization of species-dependent optical signatures of bacterial colonies by digital holography. PLoS One. 2016;11(3):e0150449.

    Article  PubMed  PubMed Central  Google Scholar 

  196. Song C, Chen Z, Zheng X, Yang S, Duan X, Jiang Y, et al. Growth characteristic analysis of Haematococcus pluvialis in a microfluidic chip using digital in-line holographic flow cytometry. Anal Chem. 2022;94(15):5769–75.

    Article  CAS  PubMed  Google Scholar 

  197. Go T, Byeon H, Lee SJ. Label-free sensor for automatic identification of erythrocytes using digital in-line holographic microscopy and machine learning. Biosens Bioelectron. 2018;103:12–8.

    Article  CAS  PubMed  Google Scholar 

  198. Go T, Kim JH, Byeon H, Lee SJ. Machine learning-based in-line holographic sensing of unstained malaria-infected red blood cells. J Biophotonics. 2018;11(9):e201800101.

    Article  PubMed  Google Scholar 

  199. Li Y, Cornelis B, Dusa A, Vanmeerbeeck G, Vercruysse D, Sohn E, et al. Accurate label-free 3-part leukocyte recognition with single cell lens-free imaging flow cytometry. Comput Biol Med. 2018;96:147–56.

    Article  PubMed  PubMed Central  Google Scholar 

  200. Singh DK, Ahrens CC, Li W, Vanapalli SA. Label-free, high-throughput holographic screening and enumeration of tumor cells in blood. Lab Chip. 2017;17(17):2920–32.

    Article  CAS  PubMed  Google Scholar 

  201. Chen D, Wang Z, Chen K, Zeng Q, Wang L, Xu X, et al. Classification of unlabeled cells using lensless digital holographic images and deep neural networks. Quant Imaging Med Surg. 2021;11(9):4137.

    Article  PubMed  PubMed Central  Google Scholar 

  202. Gangadhar A, Sari-Sarraf H, Vanapalli SA. Deep learning assisted holography microscopy for in-flow enumeration of tumor cells in blood. RSC Adv. 2023;13(7):4222–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Feizi A, Zhang Y, Greenbaum A, Guziak A, Luong M, Chan RYL, et al. Rapid, portable and cost-effective yeast cell viability and concentration analysis using lensfree on-chip microscopy and machine learning. Lab Chip. 2016;16(22):4350–8.

    Article  CAS  PubMed  Google Scholar 

  204. Sanborn D, He R, Feng L, Hong J. In situ biological particle analyzer based on digital inline holography. Biotechnol Bioeng. 2023;120(5):1399–410.

    Article  CAS  PubMed  Google Scholar 

  205. O’Connor T, Rawat S, Markman A, Javidi B. Automatic cell identification and visualization using digital holographic microscopy with head mounted augmented reality devices. Appl Opt. 2018;57(7):B197–204.

    Article  PubMed  Google Scholar 

  206. Wang Y, Ju P, Wang S, Su J, Zhai W, Wu C. Identification of living and dead microalgae cells with digital holography and verified in the East China Sea. Mar Pollut Bull. 2021;163:111927.

    Article  CAS  PubMed  Google Scholar 

  207. Terbe D, Orzó L, Zarándy Á. Classification of holograms with 3D-CNN. Sensors. 2022;22(21):8366.

    Article  PubMed  PubMed Central  Google Scholar 

  208. Park Y, Depeursinge C, Popescu G. Quantitative phase imaging in biomedicine. Nat Photonics. 2018;12:578–89.

    Article  CAS  Google Scholar 

  209. Cacace T, Bianco V, Ferraro P. Quantitative phase imaging trends in biomedical applications. Opt Lasers Eng. 2020;135:106188.

    Article  Google Scholar 

  210. Nguyen TL, Pradeep S, Judson-Torres RL, Reed J, Teitell MA, Zangle TA. Quantitative phase imaging: recent advances and expanding potential in biomedicine. ACS Nano. 2022;16(8):11516–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Merola F, Memmolo P, Miccio L, Savoia R, Mugnano M, Fontana A, et al. Tomographic flow cytometry by digital holography. Light Sci Appl. 2017;6(4):e16241.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Balasubramani V, Kuś A, Tu HY, Cheng CJ, Baczewska M, Krauze W, et al. Holographic tomography: techniques and biomedical applications. Appl Opt. 2021;60(10):B65–80.

    Article  PubMed  Google Scholar 

  213. Donnarumma D, Brodoline A, Alexandre D, Gross M. 4D holographic microscopy of zebrafish larvae microcirculation. Opt Express. 2016;24(23):26887–900.

    Article  PubMed  Google Scholar 

  214. Brodoline A, Rawat N, Alexandre D, Cubedo N, Gross M. 4D compressive sensing holographic microscopy imaging of small moving objects. Opt Lett. 2019;44(11):2827–30.

    Article  Google Scholar 

  215. Brodoline A, Rawat N, Alexandre D, Cubedo N, Gross M. 4D compressive sensing holographic imaging of small moving objects with multiple illuminations. Appl Opt. 2019;58(34):G127–34.

    Article  PubMed  Google Scholar 

  216. Dwapanyin GO, Chow DJ, Tan TC, Dubost NS, Morizet JM, Dunning KR, et al. Investigation of refractive index dynamics during in vitro embryo development using off-axis digital holographic microscopy. Biomed Opt Express. 2023;14(7):3327–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Li H, Chen X, Chi Z, Mann C, Razi A. Deep DIH: single-shot digital in-line holography reconstruction by deep learning. IEEE Access. 2020;8:202648–59.

    Article  Google Scholar 

  218. Hao J, Lin X, Lin Y, Song H, Chen R, Chen M, et al. Lensless phase retrieval based on deep learning used in holographic data storage. Opt Lett. 2021;46(17):4168–71.

    Article  PubMed  Google Scholar 

  219. Claus D, Iliescu D, Bryanston-Cross P. Quantitative space-bandwidth product analysis in digital holography. Appl Opt. 2011;50(34):H116–27.

    Article  PubMed  Google Scholar 

  220. Rubart M. Two-photon microscopy of cells and tissue. Circ Res. 2004;95(12):1154–66.

    Article  CAS  PubMed  Google Scholar 

  221. Tozer GM, Ameer-Beg SM, Baker J, Barber PR, Hill SA, Hodgkiss RJ, et al. Intravital imaging of tumour vascular networks using multi-photon fluorescence microscopy. Adv Drug Deliv Rev. 2005;57(1):135–52.

    Article  CAS  PubMed  Google Scholar 

  222. Heintzmann R, Huser T. Super-resolution structured illumination microscopy. Chem Rev. 2017;117(23):13890–908.

    Article  CAS  PubMed  Google Scholar 

  223. Olarte OE, Andilla J, Gualda EJ, Loza-Alvarez P. Light-sheet microscopy: a tutorial. Adv Opt Photonics. 2018;10(1):111–79.

    Article  Google Scholar 

  224. Wang Z, Millet L, Mir M, Ding H, Unarunotai S, Rogers J, et al. Spatial light interference microscopy (SLIM). Opt Express. 2011;19(2):1016–26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Preza C, Snyder DL, Conchello JA. Theoretical development and experimental evaluation of imaging models for differential-interference-contrast microscopy. JOSA A. 1999;16(9):2185–99.

    Article  CAS  PubMed  Google Scholar 

  226. Bochdansky AB, Jericho MH, Herndl GJ. Development and deployment of a point-source digital inline holographic microscope for the study of plankton and particles to a depth of 6000 m. Methods. 2013;11(1):28–40.

    Google Scholar 

  227. MacNeil L, Desai DK, Costa M, LaRoche J. Combining multi-marker metabarcoding and digital holography to describe eukaryotic plankton across the Newfoundland Shelf. Sci Rep. 2022;12(1):13078.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Corman R, Boutu W, Campalans A, Radicella P, Duarte J, Kholodtsova M, et al. Lensless microscopy platform for single cell and tissue visualization. Biomed Opt Express. 2020;11(5):2806–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Moon I, Javidi B. Three-dimensional identification of stem cells by computational holographic imaging. J R Soc Interface. 2007;4(13):305–13.

    Article  PubMed  Google Scholar 

  230. Delikoyun K, Yaman S, Yilmaz E, Sarigil O, Anil-Inevi M, Telli K, et al. HologLev: a hybrid magnetic levitation platform integrated with lensless holographic microscopy for density-based cell analysis. Acs Sens. 2021;6(6):2191–201.

    Article  CAS  PubMed  Google Scholar 

  231. Greenbaum A, Zhang Y, Feizi A, Chung PL, Luo W, Kandukuri SR, et al. Wide-field computational imaging of pathology slides using lens-free on-chip microscopy. Sci Transl Med. 2014;6(267):267ra175.

    Article  PubMed  Google Scholar 

  232. Rong L, Latychevskaia T, Chen C, Wang D, Yu Z, Zhou X, et al. Terahertz in-line digital holography of human hepatocellular carcinoma tissue. Sci Rep. 2015;5:8445.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  233. Dubois F, Yourassowsky C, Monnom O, Legros JC, Debeir O, van Ham P, et al. Digital holographic microscopy for the three-dimensional dynamic analysis of in vitro cancer cell migration. J Biomed Opt. 2006;11(5):054032.

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT, RS-2023-00218630).

Author information

Authors and Affiliations

Authors

Contributions

JK conducted literature searches and prepared tables and figures. JK and SJL drafted this manuscript. SJL supervised the preparation of the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Sang Joon Lee.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, J., Lee, S.J. Digital in-line holographic microscopy for label-free identification and tracking of biological cells. Military Med Res 11, 38 (2024). https://doi.org/10.1186/s40779-024-00541-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40779-024-00541-8

Keywords